1. Introduction
A (complex) C*-algebra is a Banach *-algebra
over
, and satisfies
, for all
. If
H is a complex Hilbert space then
, the bounded linear operators on
H, is a C*-algebra in the usual operator norm
, and involution
defined by
, for all
. If
and
are C*-algebras, a mapping
from
into
is called a C*-algebra
homomorphism (or, simply a *-homomorphism) if it is a homomorphism (that is, it is linear, multiplicative, and carries the unit of
onto that of
, if the C*-algebras have units) with the additional property that
for each
x in
. If, in addition,
is one-to-one (i.e., injective), it is described as a *-isomorphism. It is known that *-homomorphisms between C*-algebras are continuous ([
1], Theoerem 4.1.8), ([
2], Proposition 1.5.2). A representation of a C*-algebra
on a Hilbert space
H is a *-homomorphism
from
into
. If, in addition,
is injective, it is called a
faithful representation. An algebra
with the product
is called
a Jordan algebra if the product satisfies
, and
for all
. It is clear that if
is any associative algebra, then
,
, defines a bilinear, commutative product on
that satisfies the Jordan product identities. A Jordan Banach algebra is a real Jordan algebra
A equipped with a complete norm that satisfies
,
. The self-adjoint part
of
is a Jordan Banach algebra with the Jordan product
,
. A JB algebra is a Jordan Banach algebra A in which the norm satisfies the two identities:
and
, for all
. If
A has a unit element
, then it is clear that
. A closed Jordan subalgebra of the self-adjoint part
of all bounded linear operators
on a complex Hilbert space
H is called a JC-algebra. Sometimes a JB algebra is called a JC-algebra if it is isometrically isomorphic to a JC-algebra defined in this way, and hence, any JC-algebra is a JB algebra.
A positive linear functional
on a C*-algebra
with norm 1 is called
a state. Accordingly, the set
of all states of
is contained in the surface of the unit ball of the dual space
of
. It is known that
is convex and weak*-closed (i.e.,
-closed) ([
1], p. 257), and hence weak*-compact. By the Krein–Milman Theorem,
has an exteme point (a point
of a convex set
X in a locally convex space
is an exteme point of
X, if whenever
is expressed as a convex combination
, with
and
, then
). The extreme points of
are called
pure states, and
is the weak*-closure
of the convex hull
of the set
of its extreme points. A state
on a C*-algebra
is said to be
tracial if
for all
, equivalently, if
for all
([
2], Definition 5.3.18). A state
on a JC-algebra
A is
tracial if
for all
([
3], Defintion 5.19).
The structure of the state space of a C*-algebra can significantly influence the structure of the algebra itself, and provides valuable insights into the algebra’s properties and behavior. For example, the characteristic of an element in a C*-algebra being the zero element, self-adjoint, positive or normal is determind by the values of the states of the C*-algebra of this element ([
1], Theorem 4.3.4). There exists a longstanding known practice of employing the set of states of a C*-algebra as a dual object to illuminate the algebraic structure of the algebra. The exploration of state spaces within operator algebras, along with their geometric properties, holds significant interest due to their role in defining representations of the algebra. The intriguing and captivating aspect lies in how the algebraic structure of the system is intricately encoded within the geometric characteristics of its state space, and consequently, characterizing the state space of operator algebras among all convex sets is equivalent to characterizing the algebras (or their self-adjoint parts) among all ordered linear spaces. The great role of states lies in a fundamental result in the theory of C*-algebras and operator theory, called Gelfand–Naimark–Segal construction (GNS), which provides a powerful tool for representing abstract C*-algebras concretely as algebras of bounded operators on Hilbert spaces, connecting the C*-algebras to the familiar setting of operators on Hilbert spaces. The (GNS) construction asserts that if
is an involutive Banach algebra with unit (or, with bounded approximate identity, if it is not unital), then to each positive linear functional
on
, there is a complex Hilbert space
, a unit vector
in
, and a *-representation
of
on
such that
, for all
. The representation
is usually denoted by
, and is called
the cyclic representation of
induced by
. With
H a Hilbert space, and
a fixed element in
H, the equation
,
, defines a linear functional on
, which is clearly bounded by Cauchy–Schwarz inequality ([
1], Proposition 2.1.1), and
whenever
, that is,
is a positive linear functional on
with
, where
I is the identity of
. If
, then
is a state on
called
a vector state. Having the cyclic representation
induced by a state
of a C*-algebra
, the representation
, where
on the Hilbert space
is a faithful represntation of
, and each state of the C*-algebra
is a vector state
, for some unit vector
in
. Hence, each state of
has the form
. Identifying
with
, one can assume that any C*-algebra acts on some Hilbert space. Thus, states on given C*-algebras can be reconstructed as vector states using the cyclic representation spaces they induce. Undoubtedly, connecting C*-algebras to the familiar setting of operators on Hilbert spaces enables concrete calculations and analyses of abstract C*-algebras. This correspondence is crucial, and used in quantum theory, where states describe the physical properties of quantum systems, and understanding the formulation of quantum dynamics, and the analysis of physical observables in terms of operators is essential.
Product states of tensor product of two JC-algebras were studied by Jamjoom in [
4], and further advancements in the theory of tensor products involving infinite families of JC-algebras were established in [
5] analogous to well-known results in the context of C*-algebras. Our study aims to generalize the results delineated in [
4] to encompass the domain of infinite tensor product of JC-algebras as established in [
5]. Additionally, we characterize the tracial product state of the tensor product of two JC-algebras, as well as the tracial product state of infinite tensor products of JC-algebras. Before presenting our results in
Section 3, we provide essential background information in
Section 2 to assist readers in understanding the scope and purpose of the study.
2. Preliminaries
Tensor product of vector spaces or algebras is a fundamental concept in mathematics with broad importance and applications across various fields. It is a versatile mathematical tool with applications across a wide range of areas, including algebra, geometry, topology, physics, and engineering. It generalizes the concept of multilinear maps, and allows for the construction of a new vector space or algebra that captures multilinear relationships between vectors. In representation theory, particularly in the study of group representations and Lie algebras, tensor products provide a natural framework for combining and decomposing representations, leading to deep insights into the symmetries and structures of algebraic objects.
Let and Y be vector spaces over a field (in practice, or ). The tensor product of X and Y denoted by , or just , is characterized by the following universal propery: if is a -vector space and is a -bilinear map, then there exists a -linear map such that the following diagram commutes.
where
,
,
([
6], Section 4.7), ([
7], Section 4.5). A typical tensor in
has the form
, where
,
. Tensor products satisfy the associative and commutative laws, that is, if
, Y and
are vector spaces over a field
, then
, and
Y≅Y
([
6], Proposition 4.7.2), ([
7], Theorems 4.5.8 and 4.5.9).
Let
H and
K be (complex) Hilbert spaces. Then their algebraic tensor product
, as a vector space, is a pre-Hilbert space, where the inner product on
is defined by
,
,
,
. The completion of
(denoted also by
) is called
the Hilbert tensor product of
H and
K. Given C*-algebras
and
, their algebraic tensor product
is a complex involutary algebra in the usual way;
, and
for all
and
. By a representation of
, we mean a *-homomorphism
, where
H is a complex Hilbert space. A pair of representations
and
, where
H and K are complex Hilbert spaces, induces a natural representation
via
for all
. Given C*-algebras
and
, a C*-norm on
is a norm
satisfying
for all
. The completion of
with respect to
is a C*-algebra, and will be denoted by
. It is known that every C*-norm
on
is
a cross norm, that is,
for all
x in
and
y in
([
8], Corollary 11.3.10). The norm on
is defined by
for all
, is
the smallest (minimum) C*-norm on
, and the norm on
is defined by
for all
, is
the largest (maximum) C*-norm on
(see [
2], Definitions 4.4.5 and 4.4.8). It is convenient sometimes to write
and
instead of
and
. These norms are indeed C*-norms, and satisfy
for every C*-norm
on
([
2], p. 216), ([
8], Theorems 11.3.1 and 11.3.4). The fundamental property of the min C*-norm lies deeper, namely, given C*-algebras
and
,
, and a *-homomorphism
, then the natural map
defined by
,
, extends to a (C*-algebra) homomorphism
. Further, if
is injective, then
is injective. Hence, if
, then
([
2], Theorem 4.4.9 (iii) and Proposition 4.4.22). If
is a C*-algebra, then by taking a faithful representation
of
for each
, and identifying
with
in
, the minimum C*-tensor product is seen to be associative, in the sense that there is a *-isomorphim from
onto
taking
to
. As a result of this expressive property, given states
of
, respectively, a unique state
on
is described in terms of the given states.
Theorem 1 ([
8], Proposition 11.3.8)
. Let be a C*-algebra, and is a state of . Then there is a unique state ρ of such that , , .This state ρ is called a product state of , and is denoted by . If is acting on a Hilbert space , and is the vector state on arising from a unit vector , then is the vector state on arising from the unit vector , that is, for each The importance of product states of , derived from states on , respectively, is highlighted in the following theorem. This theorem establishes that the norm of an element can be expressed in terms of product states of . Using the associative property of the minimum C*-tensor product of the C*-algebras , we can take .
Theorem 2 (see [
2], Theorem 4.4.9 (ii)), ([
8], p. 847))
. Let be a C*-algebra, and is a state of . Then the minimum norm of each is given byLet A and B be JC-algebras, and let be a linear map. Then φ is called a Jordan homomorphism if it preserves the Jordan product, that is, for all . It is called faithful, if it is injective. It is known that any Jordan homomorphism between JC-algebras is continuous; further, if it is injective, then it is an isometry ([9], 3.4.2 and 3.4.3). Let A be a JC-algebra, a C*-algebra is called the universal enveloping C∗-algebra ofA if there is a faithful Jordan homomorphism such that generates as a C∗-algebra, and if is a C∗-algebra and is a Jordan homomorphism, then there is a *-homomorphism such that ([3], Proposition 4.36), ([9], Theorem 7.1.8). Such C*-algebra has a unique *-antiautomorphism Φ of period two leaving all points of fixed ([3], Proposition 4.40). The universal enveloping C∗-algebra ofA will be denoted by , and A will be identified with , so that A is assumed to generate as a C*-algebra. The reader is referred to [1,2,3,8,10] for the relevant material of C*-algebras, and to [4,5,11,12] for the properties of tensor products of JC-algebras and their universal enveloping C*-algebras. Definition 1. Let A and B be JC-algebras canonically embedded in the self-adjoint parts , of their respective universal enveloping C*-algebras , , so that, . Let be the Jordan algebra generated by in (see [2], Lemma 4.4.4 (i)). If λ is any C*-norm on , the completion of in is called the JC-tensor product of A and B with respect to λ. Note that given JC-algebras A and B, and a C*-norm λ on , it is not always true that ([12], Theorem 3.4). The necessary and sufficient conditions for this equality to hold are described in ([12], Proposition 2.2). Theorem 3 ([
12], Proposition 2.2 and Corollary 2.3)
. Let A and B be JC-algebras, and let or C*-cross-norm on . Then .To pass from finite to infinite tensor products of JC-algebras, we need the concept of a direct limit of a directed system of JC-algebras.
Definition 2. A directed system of JC-algebras is a family of JC-algebras in which the index set I is directed by a binary relation ≤, together with a family of Jordan homomorphisms between the JC-algebras, with the property that whenever , there is a Jordan homomorphism from into , and if such that , then .
The direct limit of a directed system of JC- algebras, denoted by , is a JC-algebra A with a family of Jordan homomorphisms that satisfies the following properties:
- 1.
is everywhere dense in A.
- 2.
For each , there is a Jordan homomorphism such that , whenever .
- 3.
If B is a JC-algebra, and is a family of Jordan homomorphisms satisfying (1) and (2) above, then there is a unique Jordan homorphism such that .
Theorem 4 ([
5], Theorem 2.2 and Corollary 2.3)
. Direct limits exist in the category of JC-algebras and Jordan homomorphisms for every directed system of JC- algebras, and .Recall that if is a JC-algebra canonically embedded in its universal enveloping C*-algebra , then by the associativity of the tensor product, we have . Sincewe have as the JC-alebra generated by in . Hence, if F is a finite set, and is a family of JC-algebras, then is the JC-algebra generated by in . Definition 3 ([
5])
. Let be a family of JC-algebras (not necessarily unitals), and let be the family of all finite subsets F of I. Define ≤ on by , if , . For each , let be the JC-tensor product of , and note that:- 1.
Whenever , , there is a natural isomorphism from onto , by Theorem 1.4 and the associativity of the tensor product.
- 2.
The map defined by is a Jordan homomorphism, by Theorem 3 and ([12], Proposition 1.2), where is an approximate identity of . - 3.
If , and , then .
Hence, the family with the Jordan homomorphisms constitutes a directed system of JC-algebras. The JC-direct limit, , of the system exists (cf. Theorem 4), and is a JC-algebra, called the tensor product of the infinite family of JC-algebras, and is denoted by ([5], Definition 2.6). The universal enveloping C*-algebra of the tensor product of an infinite family of JC-algebras is characterized in the following:
Theorem 5 ([
5], Theorem 2.7)
. Let be a family of JC-algebras. Then 3. The Main Results
Within this section, our aim is to extend the scope of Theorems 4 and 5 as outlined in [
5] (cf. Theorem 6) to encompass the realm of an infinite family of JC-algebras. Subsequently, we delve into providing the Jordan counterpart to Propositions 11.4.6, 11.3.2, and 11.4.7 elucidated in [
8]. This endeavor seeks to offer a comprehensive understanding of the applicability and implications of these propositions within the context of JC-algebras, thereby enriching the discourse and expanding the theoretical framework within this domain.
Theorem 6 ([
4], Theorems 4 and 6)
. Let A and B be JC-algebras, and let be states of A and B, respectively. Then there is a state ρ of such that , for all and . It should be noted that Theorem 6 can be applied to any finite family of JC-algebras. The state occuring in this theorem, denoted by , is called a product state of .
The characterization of the product state in the tensor product of an infinite family of C*-algebras is presented below.
Theorem 7 ([
8], Proposition 11.4.6)
. Let be a family of C*-algebras, a state of . Then there is a state ρ of the tensor product of the family such thatwhenever are distict elements of I, and , . The state occurring in Theorem 7, denoted by , is described as a product state of .
The Jordan analogue of Theorem 7 for JC-algebras is given in the following result:
Theorem 8. Let be a family of JC-algebras, and let be a state of , for each . Then there is a state ρ of such thatwhenever are distinct elements of I, and , . Proof. Let
be the family of all finite subsets
F of
I, and consider the directed system of JC-algebras
, where
is the JC-tensor product of
, and
,
, is the Jordan homomorphism defined by
strong limit
,
and
is an approximate identity of
. By ([
9], Theorem 7.1.8) and ([
1], Theorem 4.3.13(i)),
extends to a state
of
. Hence, by Theorem 7, there is a state
of the tensor product
of the family
such that for each
,
, and
,
, we have
Since
(cf. Theorem 5), the restriction
is a state of
satisfying
for each
in
, and
,
. That is,
is the product state
on the JC-subalgebra
of
. Note that if
,
, and if
,
, then we have
That is,
. Consequently, the map
, defined by
,
, is a linear bounded functional. Since
, for each
, we can easily see that
. So,
extends uniquely, by contiuity, to a state
on
, since
is everywhere dense in
, and hence, we have
.
The state occurring in Theorem 8, denoted by , is called a product state of. From the above argument, we note that given a product state on , the component states are uniquely determined, since . □
Theorem 9 ([
8], Propositions 11.3.2 and 11.4.7)
. (i) Let and be C*- algebras, and let ν and σ be states of and , respectively. Then the product state of is tracial if and only if ν and σ are tracial.(ii) Let be a family of C*-algebras, a state of . Then the product state of is pure if and only if each is pure.
The Jordan analogue of Theorem 9(i) for JC-algebras is given in the following result:
Theorem 10. Let ρ and ν be states on JC-algebras A and B, respectively. Then the product state of is tracial if and only if ρ and ν are tracial.
Proof. Suppose that
and
are tracial states. By ([
4], Theorems 4 and 6),
is a state of
, where
, for all
, and
. Now, for
, let
be a simple tensor, say
,
,
. Since
and
are tracial states, we have
, and
. Hence, by definition of the multiplication on
, we have
Now, let
, such that
,
, and
, where
and
. Then it is easy to see that
and
The linearity of
and
implies that
By the contiuity of
, we have
for all
, and hence the product state
is tracial.
Conversely, suppose that the product state
is tracial, and let
and
be increasing approximate identities of
A and
B, respectively ([
9], Proposition 3.5.4). Then
is an increasing approximate identity of
, and
, and
in norm, for all
([
12], Lemma 1.1). Let
. Since
is a state of
B,
(see ([
9], Lemma 3.6.3)), and since
is a product state, we have
Hence,
is a tracial state on
A. A similar argument shows that
is a tracial state on
B. Note that the above steps are straightforward if
A and
B are unital JC-algebras. □
Theorem 11. Let be a family of JC-algebras, and let be a state of , for each . Then the product state of is tracial if and only if is tracial for each .
Proof. Consider the directed system of JC-algebras
, where
is the family of all finite subsets
F of
I,
is the JC-tensor product of
, and
is a Jordan homomorphism of
into
, whenever
,
(cf. begining of the proof of Theorem 8). Suppose that
is tracial for each
. Then, by Theorem 10, the product state
is a tracial state on
, for each
. Note that
, where
is the product state of
occurring in Theorem 8. If
, then
, for some
, which implies that
. Since the product state
is tracial, by Theorem 10, we have
This implies that
is tricial on
, and hence, by contiuity,
is tracial on the norm closure
of
.
The converse is immidiate, since , for each .
The Jordan analogue of Theorem 9(ii) is given in the the following theorem, but first recall that if
A is JB-algebra, and
is the restriction map of the state space
of
onto the state space
of
A, then by the Krein–Milman theorem,
, where
are the set of pure states of
, respectively. The inverse image
of any
equals the line segment
, which degenerates to a single point if
, where
is the adjoint map of the *-antiautomorphism
of
([
10], Proposition 5.5). Consequently, if
is a pure state of JC-algebra
A, then there is a pure state
of
such that
(see also ([
8], Theorem 11.3.13), ([
13], Proposition 5.3.3)). □
Theorem 12. Let be a family of JC-algebras, and let be a state of , for each . Then the product state of is pure if and only if is pure for each .
Proof. We start (as in the begining of the proof of Theorem 8) by considering the directed system of JC-algebras , where is the family of all finite subsets F of I, and . Suppose that the product state of is pure. If is not pure for some , then , for some states of , , which implies that is not pure, since in this case . Hence, is pure for each .
Conversely, suppose that
is a pure state of
, for each
, and let
be a pure state of
such that
. Then
is a pure state of
, by Theorem 9(i). Since
(cf. Theorem 5), we have
as a pure state of
. Note that
is a pure state of
, for each
, by ([
8], Proposition 11.3.2), where
and
,
. Since
, we have
as a pure state of
, which satisfies
That is, for each
, the pure state
is the product state
of the pure states
of
. It follows that
and the proof is complete. □