1. Introduction
For two states $\omega$ and $\rho$
on a C$^{*}$
-algebra $\mathcal {R}$
, $\rho$
is regarded to be unitarily more mixed than $\omega$
if $\rho$
is contained in the weak* closure of the convex hull of the unitary orbit of $\omega$
. In [Reference Alberti1, Reference Alberti and Uhlmann2, Reference Wehrl29], Alberti, Uhlmann and Wehrl studied the notion of maximally unitarily mixed states on von Neumann algebras and such states were characterized by Alberti in [Reference Alberti1]. Recently, this topic has been revitalized in the broader context of C$^{*}$
-algebras by Archbold et al. [Reference Archbold, Robert and Tikuisis4], who proved among other things that the weak*closure of the set of maximally unitarily mixed states on a C$^{*}$
-algebra $A$
is equal to the weak* closure of the convex hull of tracial states and states that factor through simple traceless quotients of $A$
. However, the evolution of open quantum systems is not always unitary, but is described by more general completely positive (trace preserving) maps of the form $\omega \mapsto \sum \nolimits a_i\omega a_i^{*}$
, say on the predual of ${{\rm B}(\mathcal {H})}$
, so it seems worthwhile to study also a less restrictive notion of when one state is more mixed than the other. The dual of such a map is a unital completely positive map of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn1.png?pub-status=live)
on $\mathcal {R}={{\rm B}(\mathcal {H})}$. Let ${{\rm E}(A)}$
be the set of all unital completely positive maps on $A$
of the form (1.1), where $a_i\in A$
and the sums have only finitely many terms. A natural question in this context is, when a state $\rho$
on a C$^{*}$
-algebra $A$
(or a normal state on a von Neumann algebra $\mathcal {R}$
) is in the weak* closure (or the norm closure) of the set $\omega \circ {{\rm E}(A)}$
of all states of the form $\omega \circ \psi$
, where $\omega$
is a fixed state (perhaps normal in the case of von Neumann algebras) and $\psi$
runs over the set ${{\rm E}(A)}$
. In § 2, we show for normal states on a von Neumann algebra $\mathcal {R}$
that $\rho$
is in the norm closure of $\omega \circ {{\rm E}(\mathcal {R})}$
if and only if $\rho$
and $\omega$
agree on the centre of $\mathcal {R}$
. We also study the same topic for hermitian normal functionals on $\mathcal {R}$
and provide an explicit normal mapping $\psi$
in the point-weak* closure of ${{\rm E}(\mathcal {R})}$
such that $\rho =\omega \circ \psi$
. In the special case of $\mathcal {R}={{\rm B}(\mathcal {H})}$
, hermitian normal functionals are just hermitian trace class operators and maps mapping one such operator to another have been constructed by Hsu et al. in [Reference Hsu, Li-Wei Kuo and Tsai19] and by Li and Du in [Reference Li and Du21], but they do not study the question if such maps are in the closure of ${\rm E}({{\rm B}(\mathcal {H})})$
.
For a normal state $\omega$ on a von Neumann algebra $\mathcal {R}\subseteq {{\rm B}(\mathcal {H})}$
and a map $\phi$
of the form (1.1), where $a_i\in \mathcal {R}$
and the sums may have infinitely many terms (that is, $\phi$
is a quantum channel) any state of the form $\rho =\omega \circ \phi$
has the following property: if $\tilde {\omega }$
is a normal state on ${{\rm B}(\mathcal {H})}$
that extends $\omega$
, then there is a normal state $\tilde {\rho }$
on ${{\rm B}(\mathcal {H})}$
that extends $\rho$
such that $\tilde {\rho }$
and $\tilde {\omega }$
coincide on the commutant $\mathcal {R}^{\prime }$
of $\mathcal {R}$
(namely, $\tilde {\rho }=\tilde {\omega }\circ \tilde {\phi }$
, where $\tilde {\phi }$
is the map on ${{\rm B}(\mathcal {H})}$
given by the same formula as $\phi$
on $\mathcal {R}$
). This property holds in any faithful normal representation of $\mathcal {R}$
on a Hilbert space $\mathcal {H}$
. In § 2, we will see that this property characterizes states of the form $\omega \circ \phi$
, where $\phi$
runs over quantum channels on $\mathcal {R}$
.
Then, in § 3, we study the analogous topic for hermitian functionals $\rho,\, \omega$ on a unital C$^{*}$
-algebra $A$
. If $A$
has Hausdorff primitive spectrum, Theorem 3.1 shows that $\rho$
is in the weak* closure of $\omega \circ {{\rm E}(A)}$
if and only if $\omega$
and $\rho$
agree on the centre of $A$
and $\|c\rho \|\leq \|c\omega \|$
for each positive element $c$
in the centre of $A$
. If the primitive spectrum of $A$
is not Hausdorff, this characterization is not true any more, but an alternative one is given in Theorem 3.7.
For two states $\omega$ and $\rho$
on a C$^{*}$
-algebra $A$
, $\rho$
is regarded here to be more mixed than $\omega$
. if $\rho$
is contained in the weak* closure $\overline {\omega \circ {{\rm E}(A)}}$
of the set $\omega \circ {{\rm E}(A)}:=\{\omega \circ \psi :\, \psi \in {{\rm E}(A)}\}$
. Then, $\omega$
is called maximally mixed if for each state $\rho$
on $A$
the condition that $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
implies that $\omega \in \overline {\rho \circ {{\rm E}(A)}}$
; in other words, $\overline {\omega \circ {{\rm E}(A)}}$
is minimal among weak* closed ${{\rm E}(A)}$
-invariant subsets of the set $S(A)$
of all states on $A$
. This is a coarser relation than the one considered in the references mentioned above, where instead of ${{\rm E}(A)}$
, only convex combinations of unitary similarities are considered. In § 4, we show that each maximally mixed state on a unital C$^{*}$
-algebra $A$
must annihilate the strong radical $J_A$
of $A$
(= the intersection of all two-sided maximal ideals of $A$
) and, if $A$
is a properly infinite von Neumann algebra, the converse is also true. Furthermore, the set $S_m(A)$
of all maximally mixed states contains all states that annihilate some intersection of finitely many maximal ideals of $A$
and is therefore weak* dense in $S(A/J_A)$
. These results are analogous to those of [Reference Archbold, Robert and Tikuisis4] and [Reference Alberti1] for unitarily maximally mixed states. For C$^{*}$
-algebras with the Dixmier property, the authors of [Reference Archbold, Robert and Tikuisis4] provided a more precise determination of maximally unitarily mixed states than for general C$^{*}$
-algebras. In our present context, the role of C$^{*}$
-algebras with the Dixmier property can be played by weakly central C$^{*}$
-algebras. For a weakly central C$^{*}$
-algebra $A$
, we show that the set $S_m(A)$
is weak* closed (and hence equal to the set of all states that annihilate $J_A$
) if and only if each primitive ideal of $A$
which contains $J_A$
is maximal. States in $S_m(\mathcal {R})$
for a general von Neumann algebra $\mathcal {R}$
are also characterized.
Throughout the paper, an ideal means a norm closed two-sided ideal and all C $^{*}$-algebras are assumed to be unital unless explicitly stated otherwise.
2. The case of normal states on a von Neumann algebra
We denote by $A^{\sharp }$ the dual of a Banach case $A$
. In what follows $A$
will usually be a C$^{*}$
-algebra. Throughout this article, $\mathcal {R}$
is a von Neumann algebra, $\mathcal {R}_{\sharp }$
its predual (that is, the space of all weak* continuous linear functionals on $\mathcal {R}$
) and $\mathcal {Z}$
the centre of $R$
. Basic facts concerning von Neumann algebras, that will be used here without explicitly mentioning a reference, can be found in [Reference Kadison and Ringrose20, Reference Takesaki28].
We will need a preliminary result of independent interest, which in the special case (when, in the notation of Theorem 2.1, $\mathcal {A}=\mathcal {R}$ and $\mathcal {R}$
is a factor or has a separable predual, and positivity was not considered), has been proved by Chatterjee and Smith [Reference Chatterjee and Smith9]. We would like to avoid the separability assumption. In its proof, we will use the notion of the minimal C$^{*}$
-tensor product over $\mathcal {Z}$
of two C$^{*}$
-algebras $A$
and $B$
both containing an abelian W$^{*}$
-algebra $\mathcal {Z}$
in their centres. This product $A\otimes _{\mathcal {Z}}B$
[Reference Blanchard6, Reference Giordano and Mingo13, Reference Magajna22], can be defined as the closure of the image of the algebraic tensor product $A\odot _{\mathcal {Z}}B$
in $\oplus _{t\in \Delta }A(t)\otimes B(t)$
, where $\Delta$
is the maximal ideal space of $\mathcal {Z}$
and, for each $t\in \Delta$
, $A(t)$
denotes the quotient C$^{*}$
-algebra $A/(tA)$
, where $tA$
is the closed ideal in $A$
generated by $t$
(and similarly for $B(t)$
). (If at least one of the algebras $A$
, $B$
is exact, which will be the case in our application in the proof of Theorem 2.3, $A\otimes _{\mathcal {Z}}B$
coincides with the quotient of $A\otimes B$
by the closed ideal generated by all elements of the form $az\otimes b-a\otimes zb$
($a\in A$
, $b\in B$
, $z\in \mathcal {Z}$
) [Reference Magajna22, 3.12].)
Theorem 2.1 Let $\mathcal {A}$ be an injective von Neumann subalgebra of a von Neumann algebra $\mathcal {R}$
containing the centre $\mathcal {Z}$
of $\mathcal {R}$
. Then, each completely contractive $\mathcal {Z}$
-module map $\psi :\mathcal {R}\to \mathcal {A}$
is (as a map into $\mathcal {R}$
) in the point-weak* closure of the set consisting of all maps of the form $x\mapsto \sum \nolimits _{i=1}^{n}a_i^{*}xb_i$
($x\in \mathcal {R}$
), where $n\in \mathbb {N}$
and $a_i,\, b_i\in \mathcal {R}$
satisfy $\sum \nolimits _{i=1}^{n}a_i^{*}a_i\leq 1$
and $\sum \nolimits _{i=1}^{n}b_i^{*}b_i\leq 1$
. If in addition $\psi$
is unital, then $\psi$
is in the point-weak* closure of ${{\rm E}(\mathcal {R})}$
.
Proof. Let $\mathcal {H}$ be a Hilbert space such that $\mathcal {R}\subseteq {{\rm B}(\mathcal {H})}$
. It follows from [Reference Kadison and Ringrose20, 5.5.4] that there is a natural $*$
-isomorphism $\iota$
from $\mathcal {R}\mathcal {R}^{\prime }$
(the subalgebra of ${{\rm B}(\mathcal {H})}$
generated by $\mathcal {R}\cup \mathcal {R}^{\prime }$
) onto the algebraic tensor product $\mathcal {R}\odot _{\mathcal {Z}}\mathcal {R}^{\prime }$
, given by $rr^{\prime }\mapsto r\otimes _{\mathcal {Z}} r^{\prime }$
. By [Reference Blanchard6, 2.9], the tensor norm on $\mathcal {R}{\otimes }_{\mathcal {Z}}\mathcal {R}^{\prime }$
restricted to $\mathcal {R}\odot _{\mathcal {Z}}\mathcal {R}^{\prime }$
is minimal among all C$^{*}$
-tensor norms on $\mathcal {R}\odot _{\mathcal {Z}}\mathcal {R}^{\prime }$
, hence the $*$
-homomorphism $\iota$
extends uniquely to the norm closure $\overline {\overline {\mathcal {R}\mathcal {R}^{\prime }}}$
. Since $\mathcal {A}$
is injective and commutes with $\mathcal {R}^{\prime }$
the multiplication $\mu _0:\mathcal {A}\otimes \mathcal {R}^{\prime }\to \overline {\overline {\mathcal {A}\mathcal {R}^{\prime }}}\subseteq {{\rm B}(\mathcal {H})}$
is a completely contractive $*$
-homomorphism [Reference Brown and Ozawa8, 9.3.3, 3.8.5]. But more is true: by [Reference Giordano and Mingo13, 4.2], the natural map $\mathcal {A}\odot _{\mathcal {Z}}\mathcal {R}^{\prime }\to \mathcal {A}\mathcal {R}^{\prime }$
extends (uniquely) to a $*$
-isomorphism $\mathcal {A}\otimes _{\mathcal {Z}}\mathcal {R}^{\prime }\to \overline {\overline {\mathcal {A}\mathcal {R}^{\prime }}}$
. It follows that the composition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU1.png?pub-status=live)
is completely contractive and clearly, it is an $\mathcal {R}^{\prime }$-bimodule map, hence extending to such a map $\phi$
on ${{\rm B}(\mathcal {H})}$
by the Wittstock extension theorem (see [Reference Wittstock30] or [Reference Blecher and Le Merdy7, 3.6.2]). By [Reference Effros and Kishimoto11], $\phi$
can be approximated in the point-weak* topology by a net of elementary complete contractions of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn2.png?pub-status=live)
where $a(k)=(a_1(k),\,\ldots,\,a_n(k))^{T}$ and $b(k)=(b_1(k),\,\ldots,\,b_n(k))^{T}$
are columns with the entries $a_i(k),\,b_i(k)\in \mathcal {R}$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn3.png?pub-status=live)
Thus, $\psi$ (=$\phi |\mathcal {R}$
) can also be approximated by such maps.
Assume now in addition that $\psi$ is unital and consider a point-weak* approximation of $\psi$
of the form (2.1), (2.2). Since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU2.png?pub-status=live)
it follows that $b(k)-a(k)$ tends to $0$
in the strong operator topology. Hence, $\psi$
can be approximated by maps of the form $x\mapsto a(k)^{*}xa(k)$
in the point-weak* operator topology. To see this, write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU3.png?pub-status=live)
and note that $\|a(k)^{*}x(b(k)-a(k))\xi \|\leq \|x\|\|(b(k)-a(k))\xi \|$ for each vector $\xi \in \mathcal {H}$
. Finally, as $a(k)^{*}a(k)$
tends to $\psi (1)=1$
in the strong operator topology, $\psi$
can be approximated by maps of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU4.png?pub-status=live)
that is, by unital completely positive elementary maps.
Lemma 2.2 Let $\omega$ and $\rho$
be hermitian functionals on a C$^{*}$
-algebra $A$
such that $\rho |Z=\omega |Z$
and $\|c\rho \|\leq \|c\omega \|$
for all $c\in Z_+$
, where $Z$
is the centre of $A$
. Then, $\rho _+|Z\leq \omega _+|Z$
and $\rho _-|Z\leq \omega _-|Z$
.
Thus, if $Z$ is a von Neumann algebra, $\omega$
and $\rho$
are normal and $p^{+}$
and $p^{-}$
are the support projections of $\omega _+|Z$
and $\omega _-|Z,$
then there exist elements $c_+$
and $c_-$
in $Z$
such that $0\leq c_+\leq p^{+}$
, $0\leq c_-\leq p^{-},$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU5.png?pub-status=live)
Proof. For each $c\in Z_+$ and $\theta \in (A^{\sharp })_+$
, we have that $\|c\theta \|=(c\theta )(1)=\theta (c)$
and it is also well-known that for each hermitian functional $\sigma$
the equality $\|\sigma \|=\sigma _+(1)+\sigma _-(1)=\|\sigma _+\|+\|\sigma _-\|$
holds, hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU6.png?pub-status=live)
Adding and subtracting these two relations, we find that $\rho _+(c)\leq \omega _+(c)$ and $\rho _-(c)\leq \omega _-(c)$
for all $c\in Z_+$
. If $Z$
, $\omega$
, $\rho$
, $p^{+}$
and $p^{-}$
are as in the second part of the lemma, we may regard $Z$
as $L^{\infty }(\mu )$
for some positive measure $\mu$
and then the existence of elements $c_+$
and $c_-$
in $Z$
satisfying $0\leq c_+\leq p_+$
, $0\leq c_-\leq p^{-}$
and $\rho _+|Z=c_+\omega _+|Z$
, $\rho _-|Z=c_-\omega _-|Z$
follows easily, so we will verify here only the last equality in the lemma. The condition $\rho |Z=\omega |Z$
can be written as $(c_+\omega _+-c_-\omega _-)|Z=(\omega _+-\omega _-)|Z$
, hence $(1-c_+)\omega _+|Z=(1-c_-)\omega _-|Z$
. But $\omega _+=p^{+}\omega _+$
and $\omega _-=p^{-}\omega _-$
, since $p^{+}$
and $p^{-}$
are the support projections of $\omega _+|Z$
and $\omega _-|Z$
, hence the required equality follows.
By [Reference Halpern17] or [Reference Strătilă and Zsidó27], each positive functional $\omega$ on $\mathcal {R}$
, such that $\omega |\mathcal {Z}$
is weak* continuous, can be uniquely expressed as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn4.png?pub-status=live)
where $\omega _{\mathcal {Z}}$ is a (completely) positive $\mathcal {Z}$
-module map from $\mathcal {R}$
to $\mathcal {Z}$
such that $\omega _{\mathcal {Z}}(1)$
is the support projection $q\in \mathcal {Z}$
of $\omega |{\mathcal {Z}}$
. If $\omega$
is weak* continuous, then so is also $\omega _{\mathcal {Z}}$
. Observe that the support projections of $\omega$
and $\omega _{\mathcal {Z}}$
coincide, if $\omega$
is normal. (Indeed, for each projection $e\in \mathcal {R}$
, we have $0\leq \omega _{\mathcal {Z}}(e)\leq \omega _{\mathcal {Z}}(1)=q$
, hence $\omega (e)=(\omega |\mathcal {Z})(\omega _{\mathcal {Z}}(e))=0$
if and only if $\omega _{\mathcal {Z}}(e)=0$
since $q$
is the support projection of $\omega |\mathcal {Z}$
.)
Theorem 2.3 Let $\omega,\,\rho$ be normal hermitian functionals on $\mathcal {R}$
. There exists a normal unital completely positive map $\psi :\mathcal {R}\to \mathcal {R}$
in the point-weak* closure of ${{\rm E}(\mathcal {R})}$
satisfying $\psi (1)=1$
and $\psi _{\sharp }(\omega )=\rho$
if and only if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn5.png?pub-status=live)
Under this condition, $\rho$ is in the norm closure of $\omega \circ {{\rm E}(\mathcal {R})}$
.
Proof. Since maps in ${{\rm E}(\mathcal {R})}$ are unital and completely positive, they are also completely contractive. Each map in ${{\rm E}(\mathcal {R})}$
is of the form $\psi (x)=\sum \nolimits _{i=1}^{n}a_i^{*}xa_i$
, where $a_i\in \mathcal {R}$
and $\sum \nolimits _{i=1}^{n}a_i^{*}a_i=1$
, hence weak* continuous and the corresponding map $\psi _{\sharp }$
on the predual $\mathcal {R}_{\sharp }$
of $\mathcal {R}$
is given by $\psi _{\sharp }(\omega )=\sum \nolimits _{i=1}^{n}a_i\omega a_i^{*}$
and is a $\mathcal {Z}$
-module map with $\|\psi _{\sharp }\|=\|\psi \|=1$
. Hence $\|c\psi _{\sharp }(\omega )\|=\|\psi _{\sharp }(c\omega )\|\leq \|c\omega \|$
for each $c\in \mathcal {Z}$
. This means that the inequality $\|(c \omega )\circ \psi \|\leq \|c\omega \|$
holds for all $\psi \in {{\rm E}(\mathcal {R})}$
, hence also for all $\psi$
in the point-weak* closure of ${{\rm E}(\mathcal {R})}$
since $c\omega$
is weak* continuous. If $\psi$
is a weak* continuous such map and $\rho =\psi _{\sharp }(\omega )$
, then $\|c\rho \|=\|(c\omega )\circ \psi \|\leq \|c\omega \|$
. Furthermore, since $\psi |\mathcal {Z}={{\rm id}}$
for each such map $\psi$
, it follows that $\rho |\mathcal {Z}=\omega |\mathcal {Z}$
for each $\rho \in \overline {\omega \circ {{\rm E}(\mathcal {R})}}$
.
Assume now that the condition (2.4) holds. Decompose each of the functionals $\omega _+,\,\omega _-,\,\rho _+,\,\rho _-$ as described in (2.3), so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU7.png?pub-status=live)
where $\rho _{\mathcal {Z}}^{+},\,\rho _{\mathcal {Z}}^{-},\,\omega _{\mathcal {Z}}^{+},\,\omega _{\mathcal {Z}}^{-}$ are $\mathcal {Z}$
-module homomorphisms from $\mathcal {R}$
to $\mathcal {Z}$
such that $p^{+}:=\omega _{\mathcal {Z}}^{+}(1)$
and $p^{-}:=\omega _{\mathcal {Z}}^{-}(1)$
are the support projections of $\omega _+|\mathcal {Z}$
and $\omega _-|\mathcal {Z}$
. Let $p_+$
and $p_-$
be the support projections of $\omega _+$
and $\omega _-$
. Observe that $p_+\leq p^{+}$
and $p_-\leq p^{-}$
. (Namely, $\omega _+ (1-p^{+})=(\omega _+|\mathcal {Z})(1-p^{+})=0$
implies that $1-p^{+}\leq 1-p_+$
, hence $p_+\leq p^{+}$
.) By Lemma 2.2, there exists $c_+,\, c_-\in \mathcal {Z}$
such that $0\leq c_+\leq p^{+}$
, $0\leq c_-\leq p^{-}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn6.png?pub-status=live)
When we first tried to find a map $\psi$ satisfying the requirements of the theorem to be of the form $\psi =a\rho _{\mathcal {Z}}^{+}+b\rho _{\mathcal {Z}}^{-}$
, where $a,\,b\in \mathcal {R}_+$
, we found that it is not always possible to simultaneously satisfy the conditions $\psi (1)=1$
and $\omega \circ \psi =\rho$
by maps of such a form. But after several attempts we arrived to the following map:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn7.png?pub-status=live)
Here, $\theta$ is any fixed normal positive unital $\mathcal {Z}$
-module map from $\mathcal {R}$
to $\mathcal {Z}$
. (Such a map exists even on $\mathcal {Z}^{\prime }\supseteq \mathcal {R}$
since $\mathcal {Z}^{\prime }$
is of type $I$
, hence isomorphic to a direct sum of matrix algebras of the form ${{\rm M}}_n(\mathcal {Z})$
, where $n$
can be infinite.) This map $\psi$
is positive, weak* continuous, $\mathcal {Z}$
-module map, with the range contained in the commutative C$^{*}$
-algebra generated by $\mathcal {Z}\cup \{p_+\}$
, hence completely positive. We can immediately verify that $\psi$
is also unital:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU8.png?pub-status=live)
Now, we are going to compute
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn8.png?pub-status=live)
For this, first observe that if $f,\,g:\mathcal {R}\to \mathcal {Z}$ are $\mathcal {Z}$
-module maps and $a\in \mathcal {R}$
, then $(f\circ (ag))(x)=f(ag(x))=f(a)g(x)=(f(a)g)(x)$
, that is, $f\circ (ag)=f(a)g$
. Note also that $p^{+}\rho _{\mathcal {Z}}^{+}=\rho _{\mathcal {Z}}^{+}$
and $p^{-}\rho _{\mathcal {Z}}^{-}=\rho_{\mathcal{Z}}^{-}$
since Lemma 2.2 implies that the support projection of $\rho _+|\mathcal {Z}$
is dominated by the support projection of $\omega _+|\mathcal {Z}$
and similarly for $\rho _-|\mathcal {Z}$
and $\omega _-|\mathcal {Z}$
. From the definition (2.6) of $\psi$
and using that $\omega _{\mathcal {Z}}^{+}$
and $\omega _{\mathcal {Z}}^{-}$
are $\mathcal {Z}$
-module maps with ranges contained in $\mathcal {Z}$
and mutually orthogonal support projections $p_+$
and $p_-$
(which are just the support projections of $\omega _+$
and $\omega _-$
, respectively), we now compute
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn9.png?pub-status=live)
and similarly
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn10.png?pub-status=live)
From (2.7), (2.8) and (2.9) we have, using also (2.5) and (2.6),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU9.png?pub-status=live)
It follows from Theorem 2.1 that $\psi$ is in the point-weak* closure of ${{\rm E}(\mathcal {R})}$
. Thus, $\rho =\omega \circ \psi$
is in the weak closure of the convex set $\omega \circ {{\rm E}(\mathcal {R})}$
, which is the same as the norm closure by the Hahn–Banach theorem and the fact that $\mathcal {R}$
is the dual of $\mathcal {R}_{\sharp }$
.
When $\omega$ and $\rho$
are states, Theorem 2.3 simplifies to the following corollary:
Corollary 2.4 Let $\omega$ and $\rho$
be normal states on $\mathcal {R}$
. There exists a normal unital completely positive map $\psi$
in the point-weak* closure of ${{\rm E}(\mathcal {R})}$
satisfying $\psi _{\sharp }(\omega )=\rho$
if and only if $\rho |\mathcal {Z}=\omega |\mathcal {Z}$
. This condition is satisfied if and only if $\|c\rho \|\leq \|c\omega \|$
for all $c\in \mathcal {Z}_+$
.
Proof. By Theorem 2.3, we only need to verify that the condition $\rho |\mathcal {Z}=\omega |\mathcal {Z}$ implies that $\|c\rho \|\leq \|c\omega \|$
for all $c\in \mathcal {Z}_+$
and conversely. Since $\omega$
and $\rho$
are positive, we have $\|c\rho \|=(c\rho )(1)=\rho (c)$
and $\omega (c)=\|c\omega \|$
for all $c\in \mathcal {Z}_+$
. If $\|c\rho \|\leq \|c\omega \|$
for all $c\in \mathcal {Z}_+$
, then $\rho (c)\leq \omega (c)$
. Applying this to $1-c$
instead of $c$
, where $0\leq c\leq 1$
, it follows that $\rho (c)=\omega (c)$
for all such $c$
. But such elements span $\mathcal {Z}$
, hence it follows that $\rho |\mathcal {Z}=\omega |\mathcal {Z}$
if and only if $\|c\rho \|\leq \|c\omega \|$
for all $c\in \mathcal {Z}_+$
.
It is well known that on $\mathcal {R}={{\rm B}(\mathcal {H})}$, all normal completely positive unital maps are of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn11.png?pub-status=live)
where $\mathbb {J}$ is some set of indexes and $a_j\in \mathcal {R}$
are such that $\sum \nolimits _{j\in \mathbb {J}}a_j^{*}a_j=1$
with the convergence in the strong operator topology. Maps on ${{\rm B}(\mathcal {H})}$
of the form (2.10) are called quantum channels and we will use the same name for maps of such a form on a general von Neumann algebra $\mathcal {R}$
. It is well known that on a general von Neumann algebra, not all unital normal completely positive maps are of the form (2.10), so we still have to answer the following question: If $\omega$
and $\rho$
are normal states on a von Neumann algebra $\mathcal {R}$
, when does there exist a quantum channel $\phi$
on $\mathcal {R}$
such that $\omega \circ \phi =\rho ?$
Theorem 2.5 For normal states $\omega$ and $\rho$
on $\mathcal {R}$
, the following statements are equivalent:
(i) There exists a quantum channel $\phi$
on $\mathcal {R}$
such that $\omega \circ \phi =\rho$
.
(ii) For every faithful normal representation $\pi$
of $\mathcal {R}$
on a Hilbert space $\mathcal {H}_{\pi }$
and any normal state $\tilde {\omega }$
on ${{\rm B}(\mathcal {H}_{\pi })}$
that extends $\omega \circ \pi ^{-1}$
, there exists a normal state $\tilde {\rho }$
on ${{\rm B}(\mathcal {H}_{\pi })}$
that extends $\rho \circ \pi ^{-1}$
such that $\tilde {\omega }|\pi (\mathcal {R})^{\prime }=\tilde {\rho }|\pi (\mathcal {R})^{\prime }$
.
(iii) For some faithful normal representation of $\mathcal {R}$
on a Hilbert space $\mathcal {H}$
, such that $\omega$
is the restriction to $\mathcal {R}$
of a vector state $\tilde {\omega }$
on ${{\rm B}(\mathcal {H})}$
, there exists a normal state $\tilde {\rho }$
on ${{\rm B}(\mathcal {H})}$
such that $\tilde {\rho }|\mathcal {R}=\rho$
and $\tilde {\rho }|\mathcal {R}^{\prime }=\tilde {\omega }|\mathcal {R}^{\prime }$
.
(iv) Let $\pi _{\omega }$
be the GNS representation of $\mathcal {R}$
engendered by $\omega$
on a Hilbert space $\mathcal {H}_{\omega }$
and let $\xi _{\omega }$
be the corresponding cyclic vector. The state $\rho$
annihilates the kernel of $\pi _{\omega }$
and there exists a normal state $\tilde {\rho }$
on ${{\rm B}}(\mathcal {H}_{\omega })$
such that $\tilde {\rho }|\pi _{\omega }(\mathcal {R})$
is the state induced by $\rho$
on $\pi _{\omega }(\mathcal {R})\cong \mathcal {R}/\ker \pi _{\omega }$
and $\tilde {\rho }|\pi _{\omega }(\mathcal {R})^{\prime }=\tilde {\omega }|\pi _{\omega }(\mathcal {R})^{\prime }$
, where $\tilde {\omega }$
is the vector state $x\mapsto \langle x\xi _{\omega },\,\xi _{\omega }\rangle$
on ${{\rm B}}(\mathcal {H}_{\omega })$
.
Proof. (i)$\Rightarrow$(ii) If $\rho =\omega \circ \phi$
, where $\phi$
is of the form (2.10), then let $\tilde {\omega }$
be any state on ${{\rm B}(\mathcal {H}_{\pi })}$
extending $\omega \circ \pi ^{-1}$
, let $\tilde {\phi }$
be the map on ${{\rm B}(\mathcal {H}_{\pi })}$
defined by $\tilde {\phi }(x)=\sum \nolimits _{j\in \mathbb {J}}\pi (a_j^{*})x\pi (a_j)$
and set $\tilde {\rho }=\tilde {\omega }\circ \tilde \phi$
. Then, $\tilde {\phi }(x)=x$
for each $x\in \pi (\mathcal {R})^{\prime }$
, hence $\tilde {\rho }|\pi (\mathcal {R})^{\prime }=\tilde {\omega }|\pi (\mathcal {R})^{\prime }$
. Moreover, $\tilde {\rho }$
extends $\rho \circ \pi ^{-1}$
.
(ii)$\Rightarrow$(iii) Take for $\pi$
a faithful normal representation on a Hilbert space $\mathcal {H}$
such that $\omega$
is the restriction of a vector state $\tilde {\omega }$
on ${{\rm B}(\mathcal {H})}$
. (For example, $\mathcal {R}$
may be in the standard form [Reference Takesaki28, Chapter IX] so that all normal states on $\mathcal {R}$
and $\mathcal {R}^{\prime }$
are vector states.) For simplicity of notation, we may assume that $\mathcal {R}\subseteq {{\rm B}(\mathcal {H})}$
, that is, $\pi ={{\rm id}}$
. Then, with $\tilde {\rho }$
as in (ii), we have $\tilde {\rho }|\mathcal {R}=\rho$
and $\tilde {\rho }|\mathcal {R}^{\prime }=\tilde {\omega }|\mathcal {R}^{\prime }$
.
(iii)$\Rightarrow$(i) Assume that $\mathcal {R}$
is represented faithfully on a Hilbert space $\mathcal {H}$
such that $\omega$
is the restriction of a vector state $\tilde {\omega }$
on ${{\rm B}(\mathcal {H})}$
and that $\tilde {\rho }$
is a normal state on ${{\rm B}(\mathcal {H})}$
such that $\tilde {\rho }|\mathcal {R}=\rho$
and $\tilde {\rho }|\mathcal {R}^{\prime }=\tilde {\omega }|\mathcal {R}^{\prime }$
. Let $\xi \in \mathcal {H}$
be such that $\tilde {\omega }(x)=\langle x\xi,\,\xi \rangle$
($x\in {{\rm B}(\mathcal {H})}$
). As a normal state, $\tilde {\rho }$
is of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU10.png?pub-status=live)
where $x^{(\infty )}$ denotes the direct sum of countably many copies of $x$
acting on the direct sum $\mathcal {H}^{\infty }$
of countably many copies of $\mathcal {H}$
and $\eta \in \mathcal {H}^{\infty }$
. Now, from $\tilde {\omega }(x)=\tilde {\rho }(x)$
for all $x\in \mathcal {R}^{\prime }$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU11.png?pub-status=live)
Replacing $x$ by $x^{*}x$
, it follows that there exists an isometry $u:[\mathcal {R}^{\prime }\xi ]\to [(\mathcal {R}^{\prime })^{(\infty )}\eta ]$
such that $u\xi =\eta$
and $uy=y^{(\infty )}u$
for all $y\in \mathcal {R}^{\prime }$
. This $u$
can be extended to a partial isometry from $\mathcal {H}$
into $\mathcal {H}^{\infty }$
, denoted again by $u$
, by declaring it to be $0$
on the orthogonal complement of $[\mathcal {R}^{\prime }\xi ]$
in $\mathcal {H}$
. Then, $u$
intertwines the identity representation ${{\rm id}}$
of $\mathcal {R}^{\prime }$
and the representation ${{\rm id}}^{\infty }$
and, is therefore, a column $(u_j)$
, where $u_j\in \mathcal {R}$
. For $r\in \mathcal {R}$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU12.png?pub-status=live)
Thus, $\rho =\omega \circ \psi$, where $\psi$
is a map on $\mathcal {R}$
, defined by $\psi (r)=u^{*}r^{(\infty )}u=\sum \nolimits _ju_j^{*}ru_j$
. This map $\psi$
is not necessarily unital, but from
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU13.png?pub-status=live)
we infer that $1-u^{*}u\leq 1-p$, where $p$
is the support projection of $\omega$
. Hence, $p\leq u^{*}u$
and we may replace $\psi$
by the unital map $\phi$
defined by $\phi (r)=p\psi (r)p+(1-p)r(1-p)$
, which satisfies $\omega \circ \phi =\omega \circ \psi =\rho$
and has the required form:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU14.png?pub-status=live)
The equivalence (i)$\Leftrightarrow$(iv) is proved by similar arguments and we will omit the details, just note that $\mathcal {R}\cong \pi _{\omega }(\mathcal {R})\oplus \ker \pi _{\omega }$
.
3. The case of C$^{*}$
-algebras
In a general C$^{*}$-algebra $A$
, there are usually not enough module homomorphisms of $A$
into its centre $Z$
and even if $Z=\mathbb {C} 1$
, there can be many ideals in $A$
. Functionals on $A$
usually do not preserve ideals, hence can not be approximated by elementary operators. Therefore, we will use for general C$^{*}$
-algebras a different approach from that in the previous section, not trying to construct an explicit map sending one state to another. For C$^{*}$
-algebras with Hausdorff primitive spectrum, the situation nevertheless resembles the one for von Neumann algebras.
Theorem 3.1 Let $\omega,\,\rho$ be hermitian linear functionals on a C$^{*}$
-algebra $A$
with Hausdorff primitive spectrum $\check {A}$
and centre $Z$
. Then, $\rho$
is in the weak* closure $\overline {\omega \circ {{\rm E}(A)}}$
of the set $\omega \circ {{\rm E}(A)}$
if and only if the following condition is satisfied: (A) $\rho |Z=\omega |Z$
and $\|c\rho \|\leq \|c\omega \|$
for each $c\in Z_+$
.
Proof. To prove the non-trivial direction of the theorem, suppose that the condition (A) is satisfied, but that $\rho \notin \overline {\omega \circ {{\rm E}(A)}}$. Then, by the Hahn–Banach theorem, there exist $h\in A_h$
and $\alpha,\,\delta \in \mathbb {R}$
, $\delta >0$
, such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn12.png?pub-status=live)
Since $\rho |Z=\omega |Z$, in particular $\rho (1)=\omega (1)$
, we may replace $h$
by $h+\gamma 1$
(and $\alpha$
with $\alpha +\gamma \omega (1)$
) for a sufficiently large $\gamma \in \mathbb {R}$
and thus assume that $h$
is positive in invertible. Given $\varepsilon >0$
, let $a\in A_h$
be such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU15.png?pub-status=live)
By a well-known argument, which we now recall, this implies the relations (3.2). Namely, from the above, we have $\omega _+(1-a)<-\omega _-(1+a)+\varepsilon$ and (since $1-a\geq 0$
and $1+a\geq 0$
) this implies that $\omega _+(1-a)<\varepsilon$
and $\omega _-(1+a)\leq \varepsilon$
. Thus $\omega _+(a_+)\geq \omega _+(a)>\omega _+(1)-\varepsilon$
and $\omega _-(a_+)=\omega _-(1+a)-\omega _-(1-a_-)\leq \omega _-(1+a)\leq \varepsilon$
. In conclusion,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn13.png?pub-status=live)
For each $t\in \check {A}$ let $m(t)$
and $M(t)$
be the smallest and the largest point in the spectrum $\sigma (h(t))$
of $h(t)\in A/t$
. Since $\check {A}$
is Hausdorff by assumption, the two functions $M$
and $m$
(given by $M(t)=\|h(t)\|$
and $m(t)=\|h(t)^{-1}\|^{-1}$
) are continuous [Reference Pedersen26, 4.4.5] and therefore define elements of the centre $Z$
of $A$
by the Dauns–Hoffman theorem. Set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU16.png?pub-status=live)
For each $t\in \check {A}$, the spectrum of $b(t)$
is $\sigma (m(t)1+(M(t)-m(t))a_+(t))$
and is contained in $m(t)+(M(t)-m(t))[0,\,1]\subseteq [m(t),\,M(t)]$
since $\sigma (a_+(t))\subseteq \sigma (a_+)\subseteq [0,\,1]$
. Thus, the numerical range $W(b(t))$
of $b(t)$
(which for normal elements coincides with the convex hull of the spectrum) is contained in $W(h(t))=[m(t),\,M(t)]$
. Therefore, by [Reference Magajna23, 4.1], $b$
is in the norm closure of the set $\{\psi (h):\, \psi \in {{\rm E}(A)}\}$
, hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn14.png?pub-status=live)
by the first relation in (3.1). On the other hand, we can estimate $\omega (b)$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU17.png?pub-status=live)
(since $0\leq (M-m)(1-a_+)\leq \|M-m\|(1-a_+)$ and $0\leq (M-m)a_+\leq \|M-m\|a_+$
)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU18.png?pub-status=live)
Thus, by (3.1), (3.3) and since $m\leq h\leq M$ implies that $\rho _+(h)\leq \rho _+(M)$
and $\rho _-(h)\geq \rho _-(m)$
, we have now
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU19.png?pub-status=live)
This can be rewritten as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn15.png?pub-status=live)
or (since $\omega |Z=\rho |Z$ implies that $(\omega _--\rho _-)|Z=(\omega _+-\rho _+)|Z$
and since $M,\,m\in Z$
)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU20.png?pub-status=live)
Since this holds for all $\varepsilon >0$, by choosing small enough $\varepsilon$
, it follows that $(\omega _+-\rho _+) (M-m)<0$
. But, $Z\ni M-m\geq 0$
and $\omega _+|Z\geq \rho _+|Z$
by Lemma 2.2, hence $(\omega _+-\rho _+) (M-m)\geq 0$
, which is a contradiction.
The following corollary can be proved in the same way as Corollary 2.4, so we will omit the proof.
Corollary 3.2 If $\omega$ and $\rho$
are states on a C$^{*}$
-algebra $A$
with Hausdorff primitive spectrum, then $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
if and only if $\rho |Z=\omega |Z$
.
Before stating our main result in this section, we need a lemma. Recall that a projection $p$ in the centre of the universal von Neumann envelope $\mathcal {R}$
of a C$^{*}$
-algebra $A$
is called open if there is an ideal $J$
in $A$
such that $\overline {J}=p\mathcal {R}$
, where $\overline {J}$
is the weak* closure of $J$
in $\mathcal {R}$
.
Lemma 3.3 Let $\mathcal {R}$ be the universal von Neumann envelope of a C$^{*}$
-algebra $A$
and $\mathcal {Z}$
the centre of $\mathcal {R}$
. For each $h\in A_+$
, the central carrier $C_h$
of $h$
in $\mathcal {R}$
can be approximated in norm by linear combinations of open central projections in $\mathcal {R}$
, where the coefficients in each combination are positive.
Proof. By definition, the central carrier $z$ of $h$
is the infimum of all $c$
in $\mathcal {Z}$
such that $h\leq c$
. If $\Delta$
is the maximal ideal space of $\mathcal {Z}$
, then $z$
corresponds (via the Gelfand isomorphism) to the function $\Delta \ni t\mapsto \|h(t)\|$
, where $h(t)$
is the coset of $h$
in $\mathcal {R}/t\mathcal {R}$
. (This function is continuous by [Reference Glimm14].) Thus, we will regard $z$
as a function on $\Delta$
. Let $[m,\,M]$
be an interval containing the range of $z$
, where $m\geq 0$
and $M=\|h\|=\|z\|$
. Given $a\in A_+$
, the set $U=\{t\in \delta :\ a(t)\ne 0\}$
is open since the function $\Delta \ni t\mapsto \|a(t)\|\in \mathbb {R}$
is continuous. The weak* closure of the ideal generated by $a$
in $\mathcal {R}$
is of the form $p\mathcal {R}$
for a unique projection $p\in \mathcal {Z}$
and $p$
is open by definition. Since the quotient algebras $\mathcal {R}/t\mathcal {R}$
have only scalars in their centres, $p(t)=1$
for each $t\in U$
, hence also for each $t\in \overline {U}$
by continuity, so $p\geq q$
, where $q\in \mathcal {Z}$
is the projection that corresponds to the characteristic function of $\overline {U}$
. But from the definition of $U$
, we see that $qa=a$
and this implies that $qb=b$
for each $b$
in the ideal generated by $a$
. Hence, $qp=p$
and it follows that $q=p$
. In particular, for each $r\in \mathbb {R}_+$
the projection that corresponds to the closure of the set $U_r=\{t\in \Delta :\, z(t)>r\}$
is open since $U_r$
is just the set $\{t\in \Delta :\, a(t)\ne 0\}$
, where $a=(h-r)_+$
. (This has been observed already by Halpern in [Reference Halpern18, proof of Lemma 6].) Given $\varepsilon >0$
, for each $k\in \mathbb {N}$
let $p_k$
be the projection corresponding to the closure of the set $U_k=\{t\in \Delta :\, z(t)>M-k\varepsilon \}$
. Then, $0=p_0\leq p_1\leq p_2\leq \ldots \leq p_n=1$
, where $n\in \mathbb {N}$
is such that $M-n\varepsilon < m$
and $M-(n-1)\varepsilon \geq m$
. Now, from $1=(p_1-p_0)+(p_2-p_1)+\ldots +(p_n-p_{n-1})$
, we have that $F_k:=\overline {U}_{k}\setminus \overline {U}_{k-1}$
are disjoint closed and open sets that cover $\Delta$
and for $t\in F_k$
, we have that $M-k\varepsilon \leq z(t)\leq M-(k-1)\varepsilon$
. Thus, if we choose in each interval $[M-k\varepsilon,\,M-(k-1)\varepsilon ]$
a point $\lambda _k\geq 0$
and set $c:=\sum \nolimits _{k=1}^{n}\lambda _k(p_k-p_{k-1})$
, it follows that $\|z-c\|\leq \varepsilon$
. Finally, observe that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU21.png?pub-status=live)
is a linear combination with positive coefficients of open projections.
The following theorem is a special case of Theorem 3.7, but it is used in the proof of that theorem.
Theorem 3.4 Let $\omega$ and $\rho$
be states on a C$^{*}$
-algebra $A$
. Then, $\rho$
is in the weak* closure $\overline {\omega \circ {{\rm E}(A)}}$
of the set $\omega \circ {{\rm E}(A)}=\{\omega \circ \psi :\, \psi \in {{\rm E}(A)}\},$
where ${{\rm E}(A)}$
is the set of all unital completely positive elementary complete contractions on $A,$
if and only if $\|\rho |J\|\leq \|\omega |J\|$
for each ideal $J$
of $A$
.
Proof. Evidently, $\rho \in \overline {\omega \circ {{\rm E}(A)}}$ implies that $\|\rho |J\|\leq \|\omega |J\|$
for each ideal $J$
in $A$
since maps in ${{\rm E}(A)}$
are contractive and preserve ideals. For the converse, suppose that $\rho \notin \overline {\omega \circ {{\rm E}(A)}}$
. Then, by the Hahn–Banach theorem there exist $h\in A_h$
and $\alpha \in \mathbb {R}$
such that (3.1) holds, that is $\omega (\psi (h))\leq \alpha$
for all $\psi \in {{\rm E}(A)}$
, while $\rho (h)>\alpha.$
Replacing $h$
by $h+\beta 1$
for a sufficiently large $\beta \in \mathbb {R}$
(and consequently $\alpha$
by $\alpha +\beta$
), we may assume that $h$
is positive.
Let $\mathcal {R}$ be the universal von Neumann envelope of $A$
and denote the unique weak* continuous extensions of $\omega$
and $\rho$
to $\mathcal {R}$
by the same two letters. We will use the same notation as in the proof of Lemma 3.3. Thus, $z$
is the infimum of all $c$
in $\mathcal {Z}$
such that $h\leq c$
. Since $W(z(t)1)=\{z(t)\}\subseteq W(h(t))$
for each $t\in \Delta$
, it follows by [Reference Magajna23, 3.3] that $z\in \overline {{{\rm co}}_{\mathcal {R}}}(h)$
(= the weak* closure of the $\mathcal {R}$
-convex hull of $h$
), hence by the first relation in (3.1)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn16.png?pub-status=live)
since each map $\psi$ of the form $x\mapsto \sum \nolimits _ib_i^{*}xb_i$
($b_i\in \mathcal {R}$
, $\sum \nolimits _ib_i^{*}b_i=1$
) can be approximated by maps of the form $x\mapsto \sum \nolimits _ia_i^{*}xa_i$
($a_i\in A$
, $\sum \nolimits _ia_i^{*}a_i=1$
). (This follows by using the Kaplansky density theorem in ${{\rm M}}_n(\mathcal {R})$
to approximate the column $b=(b_1,\,\ldots,\,b_n)^{T}$
by $(a_1,\,\ldots,\,a_n)^{T}$
.) Since $\omega$
and $\rho$
are states, the hypothesis $\|\rho |J\|\leq \|\omega |J\|$
for each ideal $J$
in $A$
means that $\rho (p)\leq \omega (p)$
for each open projection $p\in \mathcal {Z}$
. Then, it follows by Lemma 3.3 that $\rho (z)\leq \omega (z)$
. But from $h\leq z$
and using (3.5), we have now that $\rho (h)\leq \rho (z)\leq \omega (z)\leq \alpha$
, which is in contradiction with the previously established relation $\rho (h)>\alpha$
.
The naive attempt to generalize Theorem 3.4 to hermitian functionals fails, as shown by the following example. The example also shows that the assumption in Theorem 3.1, that $A$ has Hausdorff primitive spectrum, is not redundant and that in Theorem 2.3 the normality of $\omega$
and $\rho$
is not redundant.
Example 3.5 For a separable Hilbert space $\mathcal {H}$, let $\omega _1$
be a normal and $\omega _2$
a singular state on ${{\rm B}(\mathcal {H})}$
, $\rho _1$
and $\rho _2$
positive normal functionals on ${{\rm B}(\mathcal {H})}$
with orthogonal supports such that $\rho _1(1)=\frac {1}{2}=\rho _2(1)$
. Set $\omega =\omega _1-\omega _2$
and $\rho =\rho _1-\rho _2$
. Then, $\rho (1)=0=\omega (1)$
, $\|\rho \|=\rho _1(1)+\rho _2(1)=1=\omega _1(1)\leq \|\omega \|$
. Since $\rho _1$
, $\rho _2$
and $\omega _1$
are normal, while $\omega _2$
is singular (which means that $\omega _2$
annihilates the ideal ${{\rm K}(\mathcal {H})}$
of all compact operators on $\mathcal {H}$
), we have $\|\rho |{{\rm K}(\mathcal {H})}\|=\|\rho \|=\rho _1(1)+\rho _2(1)=\omega _1(1)=\|\omega _1|{{\rm K}(\mathcal {H})}\|=\|\omega |{{\rm K}(\mathcal {H})}\|\leq \|\omega \|$
. Thus, $\|\rho |J\|\leq \|\omega |J\|$
for each ideal $J$
of ${{\rm B}(\mathcal {H})}$
and $\omega$
and $\rho$
agree on the centre $\mathbb {C} 1$
of ${{\rm B}(\mathcal {H})}$
. But nevertheless, $\rho \notin \overline {\omega \circ {\rm E}({{\rm B}(\mathcal {H})})}$
since on ${{\rm K}(\mathcal {H})}$
all elements of $\overline {\omega \circ {\rm E}({{\rm B}(\mathcal {H})})}$
act as elements of $\overline {\omega _1\circ {\rm E}({{\rm B}(\mathcal {H})})}|{{\rm K}(\mathcal {H})}$
and are therefore positive, while $\rho |{{\rm K}(\mathcal {H})}=(\rho _1-\rho _2)|{{\rm K}(\mathcal {H})}$
is not positive.
To generalize Theorem 3.4 to hermitian functionals, we need a lemma.
Lemma 3.6 For each hermitian functional $\omega$ on a C$^{*}$
-algebra $A$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU22.png?pub-status=live)
Proof. Suppose that $\rho \in \overline {\omega \circ {{\rm E}(A)}}$ and let $(\psi _k)$
be a net in ${{\rm E}(A)}$
such that $\rho (a)=\lim _k\omega (\psi _k(a))$
for all $a\in A$
. Extend $\omega$
, $\rho$
and each $\psi _k$
weak* continuously to the universal von Neumann envelope $\mathcal {R}$
of $A$
and denote the extensions by the same symbols. Let $\psi$
be a weak* limit point of the net $(\psi _k)$
and note that $\psi$
is a unital completely positive (hence contractive) module map over the centre $\mathcal {Z}$
of $\mathcal {R}$
. Set $\rho _1=\omega _+\circ \psi |A$
and $\rho _2=\omega _-\circ \psi |A$
. Then $\rho _1\in \overline {\omega _+\circ {{\rm E}(A)}}$
, $\rho _2\in \overline {\omega _-\circ {{\rm E}(A)}}$
and $\rho =\omega \circ \psi |A=\rho _1-\rho _2$
. This proves the inclusion $\overline {\omega \circ {{\rm E}(A)}}\subseteq \overline {\omega _+\circ {{\rm E}(A)}}-\overline {\omega _-\circ {{\rm E}(A)}}$
.
To prove the reverse inclusion, suppose that $\rho _1\in \overline {\omega _+\circ {{\rm E}(A)}}$ and $\rho _2\in \overline {\omega _-\circ {{\rm E}(A)}}$
. Then, there exist nets of maps $\phi _k$
and $\psi _k$
in ${{\rm E}(A)}$
such that $\rho _1=\lim _k\omega _+\circ \phi _k$
and $\rho _2=\lim _k\omega _-\circ \psi _k$
. Let $p$
and $q$
be the support projections in $\mathcal {R}$
of $\omega _+$
and $\omega _-$
(where $\omega _+$
and $\omega _-$
have been weak* continuously extended to $\mathcal {R}$
). Let $(a_n)$
be a net of positive contractions in $A$
strongly converging to $p$
in $\mathcal {R}$
, set $b_n=\sqrt {1-a_n^{2}}$
and define maps $\phi _{k,n}$
and $\psi _{k,n}$
on $A$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU23.png?pub-status=live)
The nets $(\omega _+(b_n^{2}))=(\omega _+(1-a_n^{2}))$ and $(\omega _-(a_n^{2}))= (\omega _-(1-b_n^{2}))$
all converge to $0$
. From this, we will verify in the following by using the Cauchy–Schwarz inequality for positive functionals that $\lim _{k,n}\omega _+\circ \phi _{k,n}=\rho _1$
, $\lim _{k,n}\omega _-\circ \psi _{k,n}=\rho _2$
, $\lim _{k,n}\omega _+\circ \psi _{k,n}=0$
and $\lim _{k,n}\omega _-\circ \phi _{k,n}=0$
pointwise on $A$
, hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU24.png?pub-status=live)
where $\theta _{k,n}:=\phi _{k,n}+\psi _{k,n}$. Evidently each $\theta _{k,n}$
is elementary completely positive map and also unital since $\theta _{k,n}(1)=a_n\phi _k(1)a_n+b_n\psi _k(1)b_n=a_n^{2}+b_n^{2}=1$
. Thus, $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
, verifying the inclusion $\overline {\omega \circ {{\rm E}(A)}}\supseteq \overline {\omega _+\circ {{\rm E}(A)}}-\overline {\omega _-\circ {{\rm E}(A)}}$
. Now, we will verify that $\lim _{k,n}\omega _+\phi _{k,n}=\rho _1$
, the verification of the other three limits that we have used is similar. For each $x\in A$
, we estimate
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU25.png?pub-status=live)
Both terms in the last line of the above expression converge to $0$.
Theorem 3.7 Let $\omega$ and $\rho$
be hermitian functionals on a C$^{*}$
-algebra $A$
. Then $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
if and only if there exist positive functionals $\rho _1$
and $\rho _2$
on $A$
satisfying the following condition:
(B) $\rho =\rho _1-\rho _2$, $\rho _1(1)=\omega _+(1)$
, $\rho _2(1)=\omega _-(1)$
, $\|\rho _1|J\|\leq \|\omega _+|J\|$
and $\|\rho _2|J\|\leq \|\omega _-|J\|$
for all ideals $J$
in $A$
.
(In particular $\|\rho |J\|\leq \|\omega |J\|$.) If $\omega$
is positive, then the condition (B) simplifies to $\rho (1)=\omega (1)$
and $\|\rho |J\|\leq \|\omega |J\|$
for all ideals $J$
.
Proof. Suppose that $\rho \in \overline {\omega \circ {{\rm E}(A)}}$. Using the notation introduced in the first part of the proof of Lemma 3.6, we have observed that the map $\psi$
on $\mathcal {R}$
introduced in that proof is a contractive unital $\mathcal {Z}$
-bimodule map. Thus, for any ideal $J$
in $A$
, if $p\in \mathcal {Z}$
is the projection satisfying $\overline {J}=p\mathcal {R}$
, then $\psi (\overline {J})=\psi (p\mathcal {R})=p\psi (\mathcal {R})\subseteq \overline {J}$
. Since $\omega$
(and hence also $\omega _+$
and $\omega _-$
) are weak* continuous on $\mathcal {R}$
, we have $\|\omega _+|\overline {J}\|=\|\omega _+|J\|$
and $\|\omega _-|\overline {J}\|=\|\omega _-|J\|$
. With $\rho _1=\omega _+\circ \psi |A$
and $\rho _2=\omega _-\circ \psi |A$
(as in the proof of Lemma 3.6), we have $\rho =\rho _1-\rho _2$
, $\rho _1(1)=\omega _+(1)$
, $\rho _2(1)=\omega _-(1)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU26.png?pub-status=live)
and similarly $\|\rho _2|J\|\leq \|\omega _-|J\|$. Therefore, also
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU27.png?pub-status=live)
Conversely, assume the existence of positive functionals $\rho _1$ and $\rho _2$
on $A$
satisfying the norm inequalities in condition (B). Then, by Theorem 3.4 $\rho _1\in \overline {\omega _+\circ {{\rm E}(A)}}$
and $\rho _2\in \overline {\omega _-\circ {{\rm E}(A)}}$
, hence by Lemma 3.6 $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
.
4. Maximally mixed states
For functionals $\omega$ and $\rho$
on a C$^{*}$
-algebra $A$
let us say that $\rho$
is more mixed than $\omega$
if $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
(where the bar denotes weak* closure). Applying Zorn's lemma to the family of all weak* closed ${{\rm E}(A)}$
-invariant subsets of $\overline {\omega \circ {{\rm E}(A)}}$
we see that in $\overline {\omega \circ {{\rm E}(A)}}$
, there exist minimal ${{\rm E}(A)}$
-invariant compact non-empty subsets, which are evidently of the form $\overline {\rho \circ {{\rm E}(A)}}$
for some $\rho$
and such $\rho$
are called maximally mixed. Thus, a functional $\omega$
is maximally mixed if $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
implies that $\omega \in \overline {\rho \circ {{\rm E}(A)}}$
. If $A$
has Hausdorff primitive spectrum, Corollary 3.2 implies that all states on $A$
are maximally mixed. The same conclusion holds for liminal C$^{*}$
-algebras.
Corollary 4.1 On a liminal C$^{*}$-algebra $A$
every state $\omega$
is maximally mixed.
Proof. If $\rho \in \overline {\omega \circ {{\rm E}(A)}}$, then by Theorem 3.4 $\|\rho |J\|\leq \|\omega |J\|$
for each ideal $J$
in $A$
. Denoting by $p$
the projection in $\mathcal {R}:=A^{\sharp \sharp }$
such that $\overline {J}=p\mathcal {R}$
, this means that $\rho (p)\leq \omega (p)$
for each open central projection $p$
, where $\omega$
and $\rho$
have been weak* continuously extended to $\mathcal {R}$
. Since $A$
is liminal, such projections are strongly dense in the set of all central projections by [Reference Digernes and Halpern10], hence it follows that $\rho (p^{\perp })\leq \omega (p^{\perp })$
. Since $\rho (p)+\rho (p^{\perp }=\rho (1)=1=\omega (1)=\omega (p)+\omega (p^{\perp })$
, we conclude that $\rho (p)=\omega (p)$
, that is $\|\rho |J\|=\|\omega |J\|$
. By Theorem 3.4, this implies that $\omega \in \overline {\rho \circ {{\rm E}(A)}}$
.
Perhaps, the simplest C$^{*}$-algebras on which not all states are maximally mixed are C$^{*}$
-algebras that have only one maximal ideal and this ideal is not $0$
.
Example 4.2 Suppose that a unital C$^{*}$-algebra $A$
has only one maximal ideal $M$
(for example, $A$
may be simple or a factor). Then, a state $\omega$
on $A$
is maximally mixed if and only if $\omega |M=0$
.
Proof. Suppose that $\omega |M=0$ and let $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
. Then, $\rho |M\!=\!0$
, hence also $\rho (J)\!=\!0$
for each proper ideal $J$
of $A$
since $J\subseteq M$
. Thus, $\|\omega |J\|=\|\rho |J\|$
for each ideal $J$
of $A$
, so $\omega \in \overline {\rho \circ {{\rm E}(A)}}$
by Theorem 3.4.
Suppose now that $\omega |M\ne 0$. Let $\rho$
be any state on $A$
such that $\rho |M=0$
. Then $\|\rho |J\|\leq \|\omega |J\|$
for all ideals $J$
, hence $\rho \in \overline {\omega \circ {{\rm E}(A)}}$
by Theorem 3.4. But $\omega \notin \overline {\rho \circ {{\rm E}(A)}}$
since $\rho |M=0$
and $\omega |M\ne 0$
, thus $\omega$
is not maximally mixed.
Remark 4.3 If $K$ is an ideal of $A$
, each state $\omega$
on $A$
satisfying $\omega (K)=0$
may be regarded as a state on $A/K$
, say $\dot \omega$
. Note that $\dot \omega$
is maximally mixed on $A/K$
if and only if $\omega$
is maximally mixed on $A$
. Indeed, denoting by $q:A\to A/K$
the natural map, $q(J)$
is an ideal in $A/K$
for each ideal $J$
in $A$
and all ideals in $A/K$
are of such a form. Moreover, $\|\omega |J\|=\|\dot \omega |q(J)\|$
, hence the claim follows from Theorem 3.4.
Example 4.2 is generalized in Theorem 4.4. The proof of Theorem 4.4 is inspired by an idea from [Reference Archbold, Robert and Tikuisis4, 3.10], but we will avoid using a background result from [Reference Archbold, Robert and Tikuisis5], that is used in [Reference Archbold, Robert and Tikuisis4, 3.10], and present a short self-contained proof. Recall that the strong radical $J_A$ of $A$
is the intersection of all maximal ideals in $A$
.
Theorem 4.4
(i) $\omega (J_A)=0$
for each maximally mixed state $\omega$
on $A$
.
(ii) If a state $\omega$
on $A$
annihilates some intersection $M_1\cap M_2\cap \ldots \cap M_n$
of finitely many maximal ideals in $A,$
then $\omega$
is maximally mixed.
Thus, the set $S_m(A)$ of maximally mixed states on $A$
is a weak* dense subset of $S(A/J_A)$
(= the set of states on $A$
that annihilate $J_A$
).
Proof.
(i) Let $D=S(A/J_A)$
and $\omega$
a maximally mixed state on $A$
. Suppose that $\omega \notin D$
. Then, $\overline {\omega \circ {{\rm E}(A)}}\cap D=\emptyset$
, otherwise this intersection would be a weak* closed proper ${{\rm E}(A)}$
-invariant subset of $\overline {\omega \circ {{\rm E}(A)}}$
, which would contradict the fact that $\omega$
is maximally mixed. Thus, by the Hahn–Banach theorem, there exist $\alpha,\,\beta \in \mathbb {R}$
and $h\in A_h$
such that
(4.1)\begin{equation} \rho(h)\leq\alpha\ \forall\rho\in D\text{ and } \omega(\psi(h))\geq\beta>\alpha\ \forall\psi\in{{\rm E}(A)}. \end{equation}Replacing $h$by $h+\gamma 1$
for a sufficiently large $\gamma \in \mathbb {R}_+$
(and modifying $\alpha,\, \beta$
), we may assume that $h$
is positive. Then, the first relation in (4.1) means that $\|\dot {h}\|\leq \alpha$
, where $\dot {h}$
denotes the coset of $h$
in $A/J_A$
. The (algebraic) numerical range $W_{A/J_A}(h)$
of $\dot {h}$
is an interval, say $[c,\,d]$
, contained in the numerical range $W_A(h)$
of $h$
, which is an interval, say $[a,\,b]$
; note that $a\leq c\leq d=\|\dot {h}\|\leq b=\|h\|$
. Let $f:[a,\,b]\to [c,\,d]$
be the function, which act as the identity on $[c,\,d]$
, and maps $[a,\,c]$
into $\{c\}$
and $[d,\,b]$
into $\{d\}$
. For every proper ideal $K$
in $A$
the quotient $A/(K+J_A)$
is non-zero, for $K$
is contained in a maximal ideal $M$
and hence $K+J_A\subseteq M+J_A=M\ne A$
. Since $W_{A/(K+J_A)}(h)\subseteq W_{A/K}(h)\cap W_{A/J_A}(h)$
, this intersection is not empty, hence the interval $W_{A/K}(h)$
intersects $[c,\,d]$
and is therefore mapped by $f$
into itself. The numerical range $W_{A/K}(f(h))$
of the coset of $f(h)$
in $A/K$
is just the convex hull of the spectrum $\sigma _{A/K}(f(h))=f(\sigma _{A/K}(h)$
, hence $W_{A/K}(f(h))\subseteq f(W_{A/K}(h))\subseteq W_{A/K}(h)$
. This inclusion implies that $f(h)\in \overline {{{\rm E}(A)}(h)}$
by [Reference Magajna23], hence $\omega (f(h))>\alpha$
by the second relation in (4.1). Since $\omega$
is a state, it follows that $W_A(f(h))$
intersects $(\alpha,\,\infty )$
. But this is a contradiction since $W_A(f(h))$
is the convex hull of the spectrum $\sigma _A(f(h))=f(\sigma _A(h))\subseteq [c,\,d]=[c,\,\|\dot {h}\|]\subseteq [c,\,\alpha ]$
. Thus, $\omega \in D$
.
(ii) By the Chinese remainder theorem [Reference Grove15, 6.3], there is a natural isomorphism $A/\cap _{j=1}^{n}M_j\cong \oplus _{j=1}^{n}A/M_j$
, thus we may regard $\omega$
as a state on $\oplus _{j=1}^{n}A/M_j$
. Since the algebras $A/M_j$
are simple, all states on them are maximally mixed by Example 4.2. The same then holds for their direct sum, so all states on $A/\cap _{j=1}^{n}M_j$
are maximally mixed and (ii) follows by Remark 4.3.
The set of all states that annihilate some finite intersection of maximal ideals of $A$ is convex and norming for $A/J_A$
(since the natural map $A/J_A\to \oplus _MA/M$
, where the sum is over all maximal ideals in $A$
, is a monomorphism, thus isometric), hence weak* dense in $S(A/J_A)$
[Reference Kadison and Ringrose20, 4.3.9].
Remark 4.5 A similar argument as in [Reference Archbold, Robert and Tikuisis4, 3.2] shows that the set $S_m(A)$ of all maximally mixed states on a C$^{*}$
-algebra $A$
is always norm closed.
Recall that a C$^{*}$-algebra $A$
is weakly central if different maximal ideals of $A$
have different intersections with the centre $Z$
of $A$
.
Theorem 4.6 If the set $S_m(A)$ of all maximally mixed states is weak* closed (which by Theorem 4.4 just means that $S_m(A)=S(A/J_A)$
), then each primitive ideal of $A$
containing $J_A$
is maximal. If $A$
is weakly central, then the converse also holds: if each primitive ideal containing $J_A$
is maximal, then $S_m(A)=S(A/J_A)$
.
Proof. By Remark 4.3 a state $\omega$ on $A/J_A$
is maximally mixed if and only if it is maximally mixed on $A$
. By [Reference Archbold and Gogić3, 3.10], the quotients of weakly central C$^{*}$
-algebras are weakly central, so in particular $A/J_A$
is weakly central. In this way, we reduce the proof to the algebra $A/J_A$
(instead of $A$
), which has strong radical $0$
. Thus, we may assume that $J_A=0$
.
Suppose now that $S_m(A)=S(A)$. Then, $S_m(A/P)=S(A/P)$
for each primitive ideal $P$
of $A$
by Remark 4.3. If $M$
is a maximal ideal of $A$
containing $P$
, then $A/M$
is a quotient of $A/P$
, hence each state $\rho \in S(A/M)$
can be regarded as a state on $A/P$
and therefore can be weak* approximated by convex combinations of vector states on $A/P$
, where $A/P$
has been faithfully represented on a Hilbert space. Since $A/P$
is primitive, as a consequence of the Kadison transitivity theorem, each vector state is of the form $x\mapsto \theta (u^{*}xu)$
for a fixed state $\theta$
on $A/P$
with $\theta (M/P)\ne 0$
, where $u\in A/P$
is unitary [Reference Kadison and Ringrose20, 5.4.5]. Thus, $\rho \in \overline {\theta \circ {{\rm E}}(A/P)}$
. But $\rho (M/P)=0$
, while $\theta (M/P)\ne 0$
if $M\ne P$
, hence $\theta \notin \overline {\rho \circ {{\rm E}}(A/P)}$
if $M\ne P$
. Thus, $\rho$
can not be maximally mixed (on $A/P$
and hence also on $A$
) if $P$
is not maximal. This argument, which we have found in [Reference Archbold, Robert and Tikuisis4, proof of 3.15], shows that in general the equality $S_m(A)=S(A)$
can hold only if all primitive ideals containing $J_A$
are maximal. If $A$
is weakly central and by our reduction above $J_A=0$
, then the assumption that all primitive ideals are maximal implies that the primitive spectrum $\check {A}$
of $A$
is homeomorphic to the maximal ideal space $\Delta$
of $Z$
(via the map $\check {A}\ni M\mapsto M\cap Z\in \Delta$
). Thus, $\check {A}$
is Hausdorff and in this case, Corollary 3.2 shows that all states on $A$
are maximally mixed.
It is well known that each W$^{*}$-algebra $\mathcal {R}$
is weakly central. If $\mathcal {R}$
is properly infinite, each primitive ideal $P$
containing $J_{\mathcal {R}}$
is maximal. (Namely, by [Reference Halpern16, 2.3] or [Reference Kadison and Ringrose20, 8.7.21], the ideal $M:=P+J_{\mathcal {R}}\supseteq \mathcal {R}(P\cap \mathcal {Z})+J_{\mathcal {R}}$
is maximal, and $M=P$
if $P\supseteq J_{\mathcal {R}}$
.) So, we can state the following corollary.
Corollary 4.7 In a properly infinite von Neumann algebra $\mathcal {R}$ maximally mixed states are just the states that annihilate the strong radical $J_{\mathcal {R}}$
.
If $\mathcal {R}$ is finite, primitive ideals are not necessarily maximal. (By [Reference Halpern17, 4.7], any ideal $\mathcal {R} t$
, where $t$
is a maximal ideal of the centre of $\mathcal {R}$
, is primitive, while using the central trace, one can show that not all such ideals are maximal in $\mathcal {R}=\oplus _{n}{\rm M}_n(\mathbb {C})$
, for example.) Thus, the set of maximally mixed states on $\mathcal {R}$
is not weak* closed.
Throughout the rest of the paper $\mathcal {R}$ is a W $^{*}$
-algebra, $\mathcal {Z}$
its centre and $\Delta$
the maximal ideal space of $\mathcal {Z}$
. For each $t\in \Delta$
let $M_t$
be the unique maximal ideal of $\mathcal {R}$
that contains $t$
[Reference Kadison and Ringrose20, 8.7.15]). Note that $\phi (\mathcal {R} t)=\phi (\mathcal {R})t\subseteq t$
for each $\mathcal {Z}$
-module map $\phi :\mathcal {R}\to \mathcal {Z}$
.
To prove that tracial states are maximally mixed, we need a lemma.
Lemma 4.8 A bounded $\mathcal {Z}$-module map $\phi :\mathcal {R}\to \mathcal {Z}\subseteq \mathcal {R}$
preserves all ideals of $\mathcal {R}$
if and only if $\phi (M_t)\subseteq t$
for each $t\in \Delta$
. If $\mathcal {R}$
is properly infinite, this is equivalent to $\phi (J_{\mathcal {R}})=0$
.
Proof. Let $J$ be an ideal in $\mathcal {R}$
and $K=J\cap \mathcal {Z}$
. As an ideal in $\mathcal {Z}$
, $K$
can be identified with the set of all continuous functions on $\Delta$
than vanish on some closed subset $\Delta _K$
of $\Delta$
, hence $K$
is the intersection of a family $\{t: t\in \Delta _K\}$
of maximal ideals of $\mathcal {Z}$
. By [Reference Kadison and Ringrose20, 8.7.15], there exists the largest ideal $J(K)$
in $\mathcal {R}$
such that $J(K)\cap \mathcal {Z}=K$
, and it follows from [Reference Kadison and Ringrose20, 8.7.16] that $J(K)=\cap _{t\in \Delta _K}M_t$
. Now $J\cap \mathcal {Z}=K$
implies that $J\subseteq J(K)$
. Thus, if $\phi$
has the property that $\phi (M_t)\subseteq t$
for all $t\in \Delta$
, then $\phi (J)\subseteq \phi (J(K))\subseteq \cap _{t\in \Delta _K}\phi (M_t)\subseteq \cap _{t\in \Delta _K}t=K\subseteq J$
.
If $\mathcal {R}$ is properly infinite, then $M_t=\mathcal {R} t+J_{\mathcal {R}}$
for each $t\in \Delta$
by [Reference Kadison and Ringrose20, 8.7.21 (1)]. Thus, if $\phi (J_{\mathcal {R}})=0$
, then we have $\phi (M_t)=\phi (\mathcal {R})t\subseteq t$
for all $t\in \Delta$
. Conversely, if $\phi (M_t)\subseteq t$
for all $t$
, then $\phi (J_{\mathcal {R}})=\phi (\cap _{t\in \Delta }M_t)\subseteq \cap _{t\in \Delta }t=0$
.
Corollary 4.9 A unital positive $\mathcal {Z}$-module map $\phi :\mathcal {R}\to \mathcal {Z}\subseteq \mathcal {R}$
is in the point-norm closure of elementary such maps (that is, $\phi \in \overline {{{\rm E}(\mathcal {R})}}^{\rm {p.n.}}$
) if and only if $\phi (M_t)\subseteq t$
for each $t\in \Delta$
.
Proof. By [Reference Magajna24, 2.2] and [Reference Magajna25, 2.1] each completely contractive map $\phi :\mathcal {R}\to \mathcal {Z}\subseteq \mathcal {R}$ which preserves all ideals of $\mathcal {R}$
is in the point-norm closure of maps of the form $x\mapsto a^{*}xb=\sum \nolimits _{j=1}^{n}a_j^{*}xb_j$
, where $n\in \mathbb {N}$
, $a_j,\,b_j\in \mathcal {R}$
, $a:=(a_1,\,\ldots,\,a_n)^{T}$
, $b:=(b_1,\,\ldots,\,b_n)$
, $\|a\|\leq 1$
and $\|b\|\leq 1$
. If $\phi$
is unital, then we can modify such maps to unital maps in the same way as in the proof of Theorem 2.1, which shows that $\phi \in \overline {{{\rm E}(\mathcal {R})}}^{{\rm p.n.}}$
.
Corollary 4.10 Let $\omega$ be a state of the form $\omega =\mu \circ \phi,$
where $\mu =\omega |\mathcal {Z}$
and $\phi :\mathcal {R}\to \mathcal {Z}$
is a unital positive $\mathcal {Z}$
-module map. If $\phi (M_t)\subseteq t$
for each $t\in \Delta,$
then $\omega$
is maximally mixed. In particular, tracial states are maximally mixed.
Proof. Suppose that $\rho \in \overline {\omega \circ {{\rm E}(\mathcal {R})}}$. Then, $\rho |\mathcal {Z}=\omega |\mathcal {Z}=\mu$
, hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU28.png?pub-status=live)
By Corollary 4.9 $\phi$ can be approximated in the point-norm topology by a net of maps $\phi _k\in \overline {{{\rm E}(\mathcal {R})}}^{{\rm p.n.}}$
. Then, $\omega (x)=\lim _k(\rho (\phi _k(x)))$
for all $x\in \mathcal {R}$
. This shows that $\overline {\omega \in \rho \circ {{\rm E}(\mathcal {R})}}$
, so $\omega$
is maximally mixed.
Any tracial state $\omega$ annihilates the properly infinite part of $\mathcal {R}$
, hence we assume that $\mathcal {R}$
is finite. Then, $\omega =(\omega |\mathcal {Z})\circ \tau$
, where $\tau$
is the central trace on $\mathcal {R}$
[Reference Kadison and Ringrose20, 8.3.10]. Since $M_t$
is of the form $M_t=\{a\in \mathcal {R}:\, \tau (a^{*}a)\in t\}$
by [Reference Kadison and Ringrose20, 8.7.17], for $a\in M_t$
, we have by the Schwarz inequality $\tau (a)^{*}\tau (a)\leq \tau (a^{*}a)\in t$
. This implies that $\tau (a)\in t$
. Thus, $\tau (M_t)\subseteq t$
, hence $\omega$
is maximally mixed by the first part of the corollary.
Are all maximally mixed states on W$^{*}$-algebras of the form specified in Corollary 4.10? Not quite. To investigate this, we still need some preparation.
Lemma 4.11 For each state $\omega$ on $\mathcal {R}$
there exists a positive $\mathcal {Z}$
-module map $\phi :\mathcal {R}\to \mathcal {Z}$
such that $\omega =(\omega |\mathcal {Z})\circ \phi$
and $p:=\phi (1)$
is a projection with $\omega (p)=1$
.
Proof. Let $\Phi$ be the universal representation of $\mathcal {R}$
, so that $\mathcal {R}^{\sharp \sharp }$
is the weak* closure of $\Phi (\mathcal {R})$
. Then, the $*$
-homomorphism $\Phi ^{-1}:\Phi (\mathcal {R})\to \mathcal {R}$
can be weak* continuously extended to a $*$
-homomorphism $\Psi :\mathcal {R}^{\sharp \sharp }\to \mathcal {R}$
; set $\tilde {\omega }=\omega \circ \Psi$
[Reference Kadison and Ringrose20, 10.1.1, 10.1.12]. Let $\tilde {\mathcal {Z}}$
be the centre of $\mathcal {R}^{\sharp \sharp }$
. Since $\tilde {\omega }$
is weak* continuous, by [Reference Halpern17] or [Reference Strătilă and Zsidó27, 1.4], there exists a unique $\tilde {\mathcal {Z}}$
-module homomorphism $\psi :\mathcal {R}^{\sharp \sharp }\to \tilde {\mathcal {Z}}$
such that $\tilde {\omega }=(\tilde {\omega }|\tilde {\mathcal {Z}})\circ \psi$
and $\psi (1)$
is the support projection $q$
of $\tilde {\omega }|\tilde {\mathcal {Z}}$
. It is not hard to verify that $\phi :=(\Psi |\tilde {\mathcal {Z}})\circ \psi \circ \Phi$
has the properties stated in the lemma.
Let $\omega$ be a state on $\mathcal {R}$
, $\mu =\omega |\mathcal {Z}$
and let $\phi$
, $p$
be as in Lemma 4.11, so that $\omega =\mu \circ \phi$
. Let $J$
be an ideal of $\mathcal {R}$
and $K=J\cap \mathcal {Z}$
. Let $(e_k)$
and $(f_l)$
be approximate units in $J$
and $K$
(respectively). Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn18.png?pub-status=live)
We may regard $(f_l)$ and $(\phi (e_k))$
as two bounded increasing nets in the positive part of the unit ball of $C(\Delta )$
($\cong \mathcal {Z}$
), hence they converge pointwise to some lower semi-continuous functions $f$
and $g$
(respectively) on $\Delta$
. The ideal $K$
of $C(\Delta )$
is of the form $K=\{a\in C(\Delta ):\, a|\Delta _K^{c}=0\}$
for some open subset $\Delta _K$
of $\Delta$
and since $(f_l)$
is an approximate unit for $K$
, it follows that $f$
is just the indicator function $\chi _{\Delta _K}$
of $\Delta _K$
. Let $\Delta _p$
be the clopen subset of $\Delta$
that correspond to the projection $p=\phi (1)$
(that is, $p=\chi _{\Delta _p}$
, the indicator function of $\Delta _p$
). Since $f_l\in J$
and $(e_k)$
is an approximate unit for $J$
, $\lim _ke_kf_l=f_l$
, hence $gf_l=\lim _k\phi (e_k)f_l=\lim _k\phi (e_kf_l)=\phi (f_l)=f_l\phi (1)=f_lp$
and $gf=\lim _lgf_l=\lim _lf_lp=fp$
, that is $(g-\chi _{\Delta _p})\chi _{\Delta _K}=0$
. This means that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn19.png?pub-status=live)
Since $(e_k)$ is an approximate unit, for any $k_1$
and $k_2$
, there exists $k_3\geq k_1,\,k_2$
so that $e_{k_3}\geq e_{k_1}$
and $e_{k_3}\geq e_{k_2}$
, and $(f_l)$
have the analogous property. Thus, $f=\sup _lf_l$
, $g=\sup _k\phi (e_k)$
and we may apply the version of the monotone convergence theorem for nets [Reference Folland12, 7.12]. Thus, denoting by $\hat {\mu }$
, the Radon measure on $\Delta$
that corresponds to $\mu$
, we have $\lim _l\mu (f_l)=\sup _l\mu (f_l)=\sup _l\int _{\Delta }f_l\,d\hat {\mu }=\int _{\Delta }\sup _l f_l\,d\hat \mu =\int _{\Delta }f\, d\hat {\mu }=\hat {\mu }(f)$
and similarly $\lim _k\mu (\phi (e_k))=\hat {\mu }(g)$
. Therefore, by (4.2), the equality $\|\omega |J\|=\|\omega |K\|$
is equivalent to $\hat \mu (g)=\hat \mu (f)=\hat \mu (\Delta _K)$
. By (4.3), this condition $\hat \mu (g)=\hat \mu (f))$
means that $0=\hat \mu (g-f)=\int _{\Delta _K^{c}\cup \Delta _p^{c}}(g-f)\,d\hat \mu =\int _{\Delta _K^{c}}(g-\chi _{\Delta _K})\,d\hat \mu =\int _{\Delta _K^{c}}g\, d\hat \mu$
, since $\hat \mu (\Delta _p^{c})=0$
(because $\mu (p)=1$
). As $g\geq 0$
, we conclude that $\|\omega |J\|=\|\omega |K\|$
if and only if $g(t)=0$
for $\hat \mu$
-almost all $t\in \Delta _K^{c}$
. Since $(e_k)$
is an approximate unit of $J$
, $\phi (e_k)(t)>0$
for some $k$
if and only if $\phi (a)(t)\ne 0$
for some $a\in J$
. Hence, since $g=\sup _k\phi (e_k)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqnU29.png?pub-status=live)
This proves the following lemma. (Note that $g$ is lower semi-continuous, hence the set $\Delta _{\phi (J)|\Delta _K^{c}\ne 0}$
in the lemma is $\hat \mu$
-measurable.)
Lemma 4.12 $\|\omega |J\|=\|\omega |(J\cap \mathcal {Z})\|$ if and only if $\hat \mu (\Delta _{\phi (J)|\Delta _K^{c}\ne 0})=0,$
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20221017231535417-0715:S0013091522000256:S0013091522000256_eqn20.png?pub-status=live)
Here, $K=J\cap \mathcal {Z}$ and $\Delta _K^{c}$
is the set of all common zeros of elements of $K$
.
The following theorem says that maximally mixed states are those for which the corresponding $\phi$ almost (with respect to $\hat \mu$
) preserve ideals.
Theorem 4.13 Let $\omega$ be any state on $\mathcal {R}$
. Let $\omega =\mu \circ \phi,$
where $\mu =\omega |\mathcal {Z}$
and $\phi :\mathcal {R}\to \mathcal {Z}$
is a positive $\mathcal {Z}$
-module map with $\phi (1)$
a projection. Denote by $\hat \mu$
the Radon measure on $\Delta$
that corresponds to $\mu$
. Then, $\omega$
is maximally mixed if and only if $\hat \mu (\Delta _{\phi (J)|\Delta _K^{c}\ne 0})=0$
for each ideal $J$
in $\mathcal {R}$
, where $K=J\cap \mathcal {Z}$
, $\Delta _K^{c}=\{t\in \Delta :\, K\subseteq t\}$
and $\Delta _{\phi (J)|\Delta _K^{c}\ne 0}$
is the set defined in (4.4).
Proof. Suppose that $\rho \in \overline {\omega \circ {{\rm E}(\mathcal {R})}}$. Then, $\rho |\mathcal {Z}=\omega |\mathcal {Z}$
and by Theorem 3.4 $\|\rho |J\|\leq \|\omega |J\|$
for each ideal $J$
of $\mathcal {R}$
. If $\hat \mu (\Delta _{\phi (J)|\Delta _K^{c}\ne 0})=0$
for each $J$
, then by Lemma 4.12 $\|\omega |J\|=\|\omega |(J\cap \mathcal {Z})\|$
for each $J$
, hence $\|\omega |J\|=\|\omega |(J\cap \mathcal {Z})\|=\|\rho |(J\cap \mathcal {Z})\|\leq \|\rho |J\|.$
Therefore, by Theorem 3.4 $\omega \in \overline {\rho \circ {{\rm E}(\mathcal {R})}}$
, which proves that $\omega$
is maximally mixed.
Conversely, if $\hat \mu (\Delta _{\phi (J_0)|\Delta _K^{c}\ne 0})>0$ for some ideal $J_0$
, then by Lemma 4.12 $\|\omega |(J_0\cap \mathcal {Z})\|<\|\omega |J_0\|$
. Let $\psi :\mathcal {R}\to \mathcal {Z}$
be any positive unital $\mathcal {Z}$
-module map that preserves ideals. (For example, the central trace, if $\mathcal {R}$
is finite, as we have seen in the proof of Corollary 4.10. If $\mathcal {R}$
is properly infinite, preservation of ideals is equivalent to $\psi (J_{\mathcal {R}})=0$
by Lemma 4.8, so we can take for $\psi$
the composition $\mathcal {R}\stackrel {\eta }{\to }\mathcal {R}/J_{\mathcal {R}}\stackrel {\iota }{\to }\mathcal {Z}$
, where $\eta$
is the natural map and $\iota$
is an extension of the inclusion $\mathcal {Z}\to \mathcal {R}/J_{\mathcal {R}}$
. Here $\mathcal {Z}$
is regarded as contained in $\mathcal {R}/J_{\mathcal {R}}$
since $\mathcal {Z}\cap J_{\mathcal {R}}=0$
, and $\iota$
exists by the C$^{*}$
-injectivity of $\mathcal {Z}$
.) Let $\rho =\mu \circ \psi$
. Since $\psi (J)\subseteq J\cap \mathcal {Z}$
for each $J$
, the set $\Delta _{\psi (J)|\Delta _{J\cap \mathcal {Z}}^{c}\ne 0}$
is empty, hence by Lemma 4.12 $\|\rho |J\|=\|\rho |(J\cap \mathcal {Z})\|$
. Since $\rho |\mathcal {Z}=\mu =\omega |\mathcal {Z}$
, we have $\|\rho |J\|=\|\rho |(J\cap \mathcal {Z})\|=\|\omega |(J\cap \mathcal {Z})\|\leq \|\omega |J\|$
for all $J$
, hence $\rho \in \overline {\omega \circ {{\rm E}(\mathcal {R})}}$
by Theorem 3.4. But $\|\omega |J_0\|>\|\omega |(J_0\cap \mathcal {Z})\|=\|\rho |(J_0\cap \mathcal {Z})\|=\|\rho |J_0\|$
implies that $\omega \notin \overline {\rho \circ {{\rm E}(\mathcal {R})}}$
. Hence, $\omega$
is not maximally mixed.
Acknowledgements
The author acknowledges the financial support from the Slovenian Research Agency (research core funding no. P1-0288).