2. Embeddings of Normed*-Algebras
In this article, infinite matrices are considered over an infinite field
F supplied with a multiplicative non-trivial norm denoted by
, where
satisfies the strong triangle inequality:
for each
x and
y in
F. It is assumed that the field
F is non-discrete and
.
Henceforth, the field F is supposed to be complete relative to its norm, if some other will not be specified.
Matrices with matrix elements belonging to F are naturally related with linear operators in normed spaces over the field F. Suppose that is a Banach space consisting of all vectors subjected to the condition
for each ,
where V is supplied with the norm
,
where is a set. For two normed spaces X and Y over the field F the linear space of all linear continuous operators is also normed:
.
Let , , , where and are sets. Then, to the operator D, a unique matrix corresponds such that and with for each , , , . The matrix is infinite, if or . Therefore, to any F-bimodule S contained in , there corresponds a F-bimodule of matrices. In particular, is an algebra of matrices over F, if S is a subalgebra in .
Assume that A is a normed algebra over the field F such that a norm on A satisfies the following conditions:
for each , also
if and only if in A,
for each and ,
and
for each a and b in A.
Frequently, it is shortly written instead of or .
We remind necessary definitions and notations.
Definition 1. Assume that the field F is of the characteristic . Assume also that is the commutative associative algebra with one generator such that and furnished with the involution for each . Suppose that A is a subalgebra in such that A is also a -bimodule, where is the Banach space over a field F, α is a set. Then, A is called a *-algebra if there is a continuous bijective (i.e., injective and surjective) F-linear operator
such that
and
and
for every a and b in A and and x and y in X, where is the canonical embedding of X into the topological dual space such that . For the sake of brevity, we can write instead of . The mapping is called the involution.
For two normed *-algebras A and B over F a map which is a continuous homomorphism of algebras and for each is called a *-homomorphism. If the *-homomorphism ϕ is bijective and is also a *-homomorphism, then ϕ is called a *-isomorphism, and the normed *-algebras A and B are called *-isomorphic.
For the normed *-algebra A and (or ) by (or , respectively) will be denoted a minimal normed subalgebra in A containing a (or U, respectively). By will be denoted the closure of in A.
Remark 1. In Definition 1, is a particular case of a bilinear functional (see [11] and Remark 2 in more details). Note that for matrices with entries in corresponding to operators on -linear spaces, their block form with entries in is frequently used, because each complex number can be written as the matrix with real entries. This is utilized for generating a complex *-algebra by the doubling procedure from a corresponding real algebra [1,3,5,26]. Using similar ideas, one can construct examples of Banach *-algebras over the field F other than and with . This is evident for with for primes p such that by Corollary 6 in Ch. 1, Section 4 in [25]. Example 1. Let be the Banach space such that . In view of Theorem 5.13 in [11], the direct sum is isomorphic with X. In more details, one can take for the set α any fixed partition with and , . Therefore, is isomorphic with X for each and is isomorphic with X. This also implies that is isomorphic with as the Banach algebra for each . For row vectors and operators , it is frequently written for convenience instead of or , where is a matrix corresponding to an operator a. Assume that . For the algebra can be embedded into in the standard way, up to an automorphism of the Banach algebra , as induced by the formula for each , where and , , I is the unit operator on X.
Example 2. Let the condition of Example 1 be satisfied. Let B be a Banach subalgebra in and let be any continuous antiautomorphism of B. That is ψ is an F-linear map with continuous ψ and , ψ is a bijection (i.e., injection and surjection), for each a and b in B. Such antiautomorphisms always exist, for example, for each a and b in B [1,5,27]. Shortly, we denote by . There are natural embeddings of into as the Banach algebras such that for each , for each , for each with , , where . Then, to each or , one can pose operators with and with for each and . Shortly, , , , will be denoted by , respectively.
Let ψ be the antiautomorphism of . The Banach algebras and are isomorphic, hence ψ on induces ψ on . For each let be such that for each and . Notice that, with as in Example 1, evidently is the -bimodule, where for subsets U and W in a F-linear space Y.
We take the antiautomorphism ψ of and extend it on such that and for each and in . For any and in , we put and . Hence, , and , since . Therefore, for each and with , , , in , since , , , . This implies that is the closed subalgebra in and is supplied with the *-algebra structure.
Example 3. Let the conditions of Example 2 be satisfied. We take any fixed antiautomorphism ψ on extended on and inducing the involution as in Example 2. Certainly, for any given subset V in , there exists a minimal closed subalgebra in such that , is the -bimodule, . This algebra is the intersection of all closed *-subalgebras in such that and is the -bimodule, where is with the involution inherited from and . Evidently, is the closure in of a family of all operators of the following form:
belonging to with in F, , ..., in , where .
Take , with , , , , where , . Therefore, is isomorphic with and with for each . In particular, V can be taken contained in , since there is the natural embedding for analogously to for described above. Therefore, there exists , such that and is the nontrivial *-algebra.
Example 4. Let the conditions of Examples 1 and 2 be satisfied. For Banach spaces B and Y over F by is denoted the completion relative to the projective tensor product topology (see [11,28]) of the tensor product over the field F; or shortly instead of , if F is specified. Let also , where . In view of Theorems 4.33, 4.40, and 4.41 in [11], and is the closed subalgebra in , where denotes the algebra of all compact operators from into . Notice that is isomorphic with by Theorem 4.41 [11], where is the topological dual space of , . We say that is the algebra of bicompact operators. In this case, it is possible to take for each , where for each j and k in . This implies that (see Example 3). Hence, . Example 5. Other examples of *-algebras which are proper subalgebras in can be provided utilizing combinations of Examples 3 and 4. Note also that finite direct sums of *-algebras are *-algebras.
Assume that Λ is an infinite set. For Banach spaces over F for each , let denote a Banach space over F such that each has the form and , , , for each , for each x and y in . Assume also that is a Banach *-algebra over the field F for each . Then, is the Banach *-algebra with multiplication and inversion for each x and y in G. We call G the direct sum of the Banach *-algebras and denote it also by . Similarly, a direct sum can be defined.
4. -Matrices and Algebras
Definition 4. Let A be an normed algebra over the field F, (see Introduction), satisfying the following conditions:
A is a Banach *-algebra and
there exists a bilinear functional such that for all x and y in A, where is a constant independent of x and y,
and for each x and y in A,
if for each , then ;
for every x, y and z in A,
for each nonzero element .
Then we call A a -algebra. If an operator D belongs to the -algebra A, , , then the corresponding matrix is called a -matrix.
Lemma 2. For a *-subalgebra A of with , a bilinear functional satisfying conditions , , and exists.
Proof. We put , where S is a marked compact operator such that , , (see Remark 2). From the inequality it follows that Condition is valid. From for each and property follows, since for each . Then, using the identity for each and we deduce that for every x, y and z in A, since . □
Lemma 3. Let and let A be a Banach *-algebra over F such that . Let also either or be satisfied:
- (i)
If and for each there exists a normed extension of the field F such that there exists a *-homomorphism from into a *-algebra , such that with and , where is an invertible operator and a matrix of an operator is diagonal and nonzero, ; or
- (ii)
if and for each there exists a *-homomorphism with nonzero image , (see Examples 2 and 3),
then conditions – are also valid. Moreover,
- (iii)
if is a -algebra over F for each ,
then and are -algebras.
Proof. In cases and , in view of Lemmas 2, 1, and Formula , Conditions , and are satisfied. Indeed, using injective *-homomorphism it is possible to choose for which the decomposition is such that is an automorphism of the Banach space X and , also with for each j, while for each , where is the standard basis of X, since for each . Then, we get property , since .
In case , we deduce that , hence , since is the *-homomorphism and . This implies .
In case , let . By the imposed conditions in x is nonzero, . On the other hand, and for each . Let , for each j and k in (see also Definition 1 and Examples 2, 3). Therefore, considering of the form with with for every , j and k in , one finds coefficients such that , since for each in , . Note that implies that and consequently, , since the algebra A is associative and is the *-homomorphism. Thus, property also is fulfilled.
. For each (or ), there is the decomposition
(or , respectively) with for each . Therefore, (or , respectively). Hence if , then .
For each there exists a constant such that for each , in , where denotes the bilinear functional on satisfying the conditions of Definition 4. We choose such that , because the field F is infinite non-discrete. For each there exists such that . Let
,
then for each x and y in A (or B, respectively). This implies that the bilinear functional given by Formula satisfies conditions of Definition 4. □
Lemma 4. If and are proper or improper right and left ideals in a -algebra A, then and are orthogonal relative to the family of bilinear functionals complements of the sets and in the Banach space A, where for every a, x and y in A.
Proof. If (see Definition 2), then , hence for each and consequently, by identity and inevitably . This means that relative to , that is is the orthogonal complement of . Similarly, is the orthogonal complement of in A as the Banach space relative to the family of bilinear functionals. □
Proposition 1. Any -algebra A is dual.
Proof. If and are right and left ideals in A, then by Lemma 4 and analogously , since and . □
Theorem 1. Any -algebra A over the spherically complete field F with is representable as the direct sum of its two-sided minimal closed ideals, which are simple -algebras and pairwise orthogonal relative to the family of bilinear functionals .
Proof. By virtue of Theorem 8 in [
9] and Proposition 1, the algebra
A is the completion (relative to the norm) of the direct sum of its minimal closed two-sided ideals which are simple dual subalgebras (see also Definition 3). Consider a two-sided minimal closed non null ideal
J in
A. The involution mapping
provides from it the minimal closed two-sided ideal
due to Condition
.
Suppose that , then , since the ideal J is minimal. From and for each we deduce that . Together with condition imposed on the -algebra, this would imply that for each contradicting . Thus, .
Notice that properties – and for J are inherited from that of A. Then, condition on A implies that , since and , also . However, J is minimal, hence . Therefore, property on J follows from that of on A and and , since for each there exists x and y in J with and for all , also since . Then, for each an element exists such that , hence . Then, we have that by on A. Hence, , consequently, , since the algebra A is associative and . Therefore, property on J is valid. Thus, J is the -algebra.
If J and S are two distinct minimal closed two-sided ideals in A, then . From Lemma 4, it follows that . Thus, these ideals J and S are orthogonal relative to the family of bilinear functionals.
Using condition and Lemma 4, we infer that A is the direct sum of its two-sided minimal closed ideals. □
Theorem 2. Let A be a simple unital -algebra over the spherically complete field F with and let a division algebra G be provided by Theorem 2 in [10]. Then, the following conditions are equivalent: - (i)
is finite dimensional over G;
- (ii)
is unital;
- (iii)
the center of is non-null.
Proof. Let
be a maximal system of irreducible idempotents provided by Theorem 2 in [
10].
. If is finite dimensional over G, then according to Theorem 1, a maximal system of irreducible idempotents is finite, that is . Then, their sum is the idempotent fulfilling the condition and . Thus, w is the unit in .
. If contains a unit w, then contains w, consequently, is non-null.
. Let
and
x be a non-zero element of
,
. In view of Theorem 2 in [
10]
, hence
, where
. Thus
. Therefore,
and hence
. Similarly,
, consequently,
and hence
.
Note that for each j, where plays the role of the unit in . Then,
for each j and k, hence is the isomorphism of normed algebras with for each j and k.
Therefore, the sum may converge only if it is finite. Thus, the algebra is finite dimensional over G. □
Remark 3. For a Banach space H over the field F and a set α by is denoted a direct sum of α copies of H such that is a Banach space consisting of all vectors with and such that for each a set is finite. In particular, for the Banach space over a spherically complete field F, there exists a topological dual space of all continuous F-linear functionals (see Ch. 2 and 5 in [11] or Ch. 8 in [29]). Each vector x in has the following decomposition: , where , with , denotes the Kronecker delta symbol such that for each in α, for each . Then, for a division algebra H over the spherically complete field F and a Banach H-bimodule we consider a bounded F-linear right H-linear operator C from into , that is for each and . The embedding of F into H as , where is a unit element in H, induces a F-linear embedding of X into . In this case to each there corresponds a continuous F-linear right H-linear functional such that for each and . This induces a natural embedding , where denotes a space of all bounded F-linear right H-linear operators from into H (see Ch. 3 and 5 in [11], Proposition 23.1 in [16]). Therefore, for the operator C and for each i and j in α, there exists a matrix element . Then, by is denoted the space of all bounded F-linear right H-linear operators C from into satisfying the condition: - (i)
for each a finite subset γ in a set α exists such that for each j and k with either or .
Theorem 3. Let A be a spherically complete simple unital -algebra over a spherically complete field F with . Let also G be a division algebra provided by Theorem 2 in [10] such that for each , also and . Then a Banach G-bimodule exist such that A and are isomorphic as the Banach right G-modules and as F-algebras. Proof. By the conditions of this theorem, a division algebra G is such that for each irreducible idempotent w in A. Put . From , it follows that . If , then and , since , consequently, .
For each irreducible idempotent
w such that
(see the proof of Theorem 3 in [
10]) one gets that
, since
A is the
-algebra over
F. Then,
, hence
and consequently,
implying that
, since
and
for each
. Therefore,
also.
Since w is the irreducible idempotent and , then is the irreducible idempotent in the -algebra A. Then, we deduce that , since , consequently, an element exists such that , since . The latter implies . However, the elements and are self-adjoint, hence and consequently,
.
We put , hence
and
.
Thus, v is the self-adjoint idempotent. On the other hand, and and the idempotent w is irreducible, hence the idempotent v is also irreducible, since is the non-null minimal left ideal in .
Then, from the proof of Theorem 2 in [
10] it follows that
is the self-adjoint division algebra for each such irreducible self-adjoint idempotent
v, consequently,
. By the conditions of this theorem we have
.
The algebra A is simple, that is by the definition each its two-sided ideal coincides with either or A.
Next we take a maximal orthogonal system
of self-adjoint idempotents in
A and for them elements
as in Theorem 2 in [
10], where
is a set. Hence,
and
exists such that
. Then,
, since
and
. Moreover,
, since
is non null and hence
is non-null. For
, we deduce that
, since
,
since A is associative and for each non null b in H, where .
Thus, it is possible to choose an element such that for each k. Taking a marked element and setting and for each l and k one gets and , also for every . Thus, elements can be chosen such that for each l and k.
If the statement of this theorem for the spherical completion
of
H is proven, then it will imply the statement of this theorem for
H. So the case of the spherically complete division algebra
H is sufficient. Then,
A and
H considered as the Banach spaces over the spherically complete field
F are isomorphic with
and
H with
due to Theorems 5.13 and 5.16 in [
11], where
.
From the proof of Theorem 3 in [
10], it follows that the sum
is dense in
A. Conditions
,
,
imply that
, since
for each
. Therefore, from properties
,
,
it follows that if
or
, then
for each
x and
z in
A, since
for each , also
for each .
Thus, the set is complete and for each or , where the latter property is interpreted as the orthogonality. Together with property , this implies that each element has the form with , since is the right H-module, also A is isomorphic with as the F-algebra and the right G-module, where the series may be infinite, for each , where denotes the corresponding set.
Take the Banach H-bimodule and to each element one can pose the operator such that (see Remark 3), where and for each j and k in , since for each , where (see above). Then, and the mapping is the isometry having the isometrical extension . The property given above provides for each j and , consequently, T is bijective from A onto , since A is simple.
For each S and V in , one has for each and . Moreover, , consequently, satisfies condition in Remark 3, that is . Hence, by verifying other properties, one gets that also has the F-algebra structure. From the construction of , it follows that is the F-algebra, since H and A are F-algebras. Notice that, moreover, as the F-algebra is isomorphic with the Banach F-algebra . By the conditions of this theorem, is isomorphic with A as the F-algebra and the right H-module. □
Theorem 4. Let A be a spherically complete simple unital -algebra over the spherically complete field F with and . Let also G be a division algebra provided by Theorem 2 in [10] such that for each . Then a division subalgebra H of G and a Banach H-bimodule exist such that and are isomorphic as the Banach right H-modules and as F-algebras. Proof. In this case, and instead of A we consider .
The
-algebra
A is simple and central,
, hence the right
H-module
is simple due to Satz 5.9 in [
7] and Theorem 2 above. We denote
shortly by
A and the rest of the proof is similar to that of Theorem 3. □
From Theorems 1, 3 and 4, the corollary follows.
Corollary 1. Suppose that A is a spherically complete unital -algebra over the spherically complete field F with and G is the division algebra given by Theorem 2 in [10] so that for each such that either and or
.
Then, a division subalgebra H in G with and H-bimodules exist such that as the right H-module and the F-algebra is the direct sum of .
Example 6. Let A be a -algebra over a spherically complete normed field F (see Definition 4 and Introduction). Evidently, the algebra A also has the structure of the Banach A-bimodule. Hence, there exists a Banach space H over F such that A can be embedded into the normed algebra of all bounded F-linear operators . In view of Theorems 5.13 and 5.16 in [11], there exists a set α such that H is isomorphic with the Banach space (see Remark 3). Therefore, each element D of A is characterized by the corresponding to it matrix , which is unique relative to a fixed basis in H. This matrix is infinite, if .