1. Introduction
In recent years, symmetric structures and mappings in infinite-dimensional spaces have been studied by numerous authors [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11]. In many problems of algebra and analysis [
1,
6], as well as in applications in symmetric neural networks (see, e.g., [
12,
13,
14,
15]), it is crucial to know the invariants of a given (semi)-group
acting on a Banach space
The invariants can be described as elements of algebras of
-symmetric functions on
The Classical Invariant Theory, which was developed in the middle of the last century, investigated polynomial invariants of a group acting on a finite-dimensional linear space. The famous Nagata counterexample to the general case of Hilbert’s fourteenth problem shows that polynomial algebras on
may be not finitely generated.
Symmetric polynomials and analytic functions on infinite-dimensional Banach spaces were investigated first by [
16,
17,
18,
19]. In particular, in [
16,
17], algebraic bases were described in algebras of symmetric polynomials on various Banach spaces with symmetric structures. These investigations were continued in [
19,
20,
21,
22,
23,
24,
25,
26] and others. To describe the spectrum of a uniform algebra of
-symmetric functions on
X, it is important to have more information about the quotient set
X/
where “∼” is the relation of equivalence “up to the action of
” on
Such a quotient set may be interesting in itself and has applications in informatics and neural networks. If
X is a sequence space and
is the group of permutations of elements of the sequences, then
X/∼ can be considered as a set of nonzero multisets—completed in a metrizable topology—induced from
The set
X/∼ has a semiring structure with respect to natural algebraic operations. The commutative semiring can be extended to a ring by using a standard procedure from
K-theory (see, e.g., [
27]). Such a ring
of multisets for the case
was investigated in [
7,
28]. In particular, homomorphisms and ideals of
were considered, and it was shown that each supersymmetric polynomial on
can be extended to the ring
In [
29], the properties of the ring of multisets of integer numbers were studied, and some applications to cryptography were found.
In this paper, we consider possible generalizations of the results obtained in [
7] for more general cases. Instead of the sequence space
we consider the space of sequences
where
are elements of a Banach algebra
and each sequence of norms,
, is a vector in a Banach space
X with a norm
and a symmetric basis
Let us recall (see [
30] for details) that a sequence
is a
topological (or
Schauder) basis in a Banach space
X if every element
can be uniquely expressed by
where the limit is taken in
From here, in particular, we have that
as
A topological basis is called
symmetric if it is equivalent to the basis
for every permutation
on the set of natural numbers
This means that for every
a series
converges if and only if
converges. It is known [
30] (p. 114) that every Banach space
X with a symmetric basis has an equivalent so-called
symmetric norm such that
for every permutation
and sequence of numbers
such that
Throughout this paper, we assume that
X is endowed with a symmetric norm. In this case, we know that for every
In
Section 2, we construct a ring of multisets
of sets from a multiplicative semigroup
of
and investigate the basic properties. In particular, we show that
is complete in a metrizable topology induced from
In
Section 3, we investigate homomorphisms of
and related supersymmetric polynomials. In addition, we consider some examples and make discussions. We refer the reader to [
31] for more information about polynomials on Banach spaces and to [
32] for details on the classical theory of symmetric functions.
2. Group Rings of Multisets
Let
X be a Banach space with a normalized symmetric basis
and a symmetric norm
let
be a Banach algebra with an identity
and let
be a closed multiplicative subgroup in
containing
We denote by
the set of sequences
and
In addition, let us denote by
, and we represent each element
as
Clearly,
is a Banach space with respect to the norm
and
is its closed subset.
For a given , we denote by the subset of all natural numbers such that
Let
be permutations on
and
We define
Let
and
be in
Then,
Note that if then Hence, for all
Let us consider an equivalence defined as
if and only if there are vectors
and bijections
and
such that
maps
onto
and
maps
onto
; in addition,
Let us denote by the quotient set with respect to the equivalence “∼”. We denote by the class of equivalence containing element Clearly, for every , and so In addition, we denote
Let us explain the definition of the equivalence in a more detailed form. The requirement that
and
act bijectively between supports of corresponding vectors means that zero coordinates “do not matter”, that is, for example,
In addition, for example,
for any
In addition, the classes of equivalence are invariant with respect to permutations of coordinates of
x and of
y separately. This approach allows us to consider
as a set of multisets of
More exactly, the subset
consisting of all elements in
with finite supports can be naturally identified with the set of all finite multisets of
nonzero elements in
, and the operation “•” is actually the union of multisets.
We say that
is an
irreducible representative of
if
, and
implies that
for some permutations
on
and
In other words, for every nonzero coordinate
of
, we have
for all coordinates
of
Proposition 1. For every , there exists an irreducible representative.
Proof. Let
be a representative of
Since elements
and
belong to the Banach space
X with the Schauder basis
it follows that
and
as
Without loss of generality, we may assume that the coordinates of
x are ordered so that
If there is
j such that
then let us remove the coordinate
in
x and
in
and we denote by
and
the resulting vectors. If such a number
j does not exist, we denote
and
Suppose that
and
are already constructed. If there is
j such that
then we remove the coordinate
in
and
in
and denote by
and
the resulting vectors. Otherwise, we set
and
Thus, we obtain the sequence
in
, which is obviously fundamental. By the completeness of
there exists a limit
Let a be a vector in such that its coordinates are exactly removed coordinates from Then, , and so is a representative of By the construction, is irreducible. □
Now, we can introduce a commutative operation “+” on
Definition 1. For a given and in , we define In addition, we set
Proposition 2. The operation “+” is well defined on and is a commutative group with zero (the neutral element),
Proof. From definition of the operation, it follows that
and
If
and
are the irreducible representatives
and
then, according to (
1) and Proposition 1,
and
for some
a and
Hence,
So, the result does not depend of representatives. □
Let By , we denote the resulting sequence of ordering the set with one single index in some fixed order.
Proposition 3. Let Then, and Moreover, if is such that for every and or for some then
Proof. Let
be a bijection from
to
According to the straightforward calculations,
Let
be such that
for every
If
then
□
Next, let us define a multiplication on
Definition 2. If and we define Finally, if and are in then we define Using routine calculations, it is easy to check (cf. [
7,
29]) that the multiplication is well defined and associative and that the distributive low with the addition holds on
If
is a commutative Banach algebra, then the introduced multiplication is commutative. So, we have the following proposition.
Proposition 4. is a ring with zero, , and unity, If is commutative, then is commutative.
Note that
is not an algebra, even if
, because it is not a linear space (see, e.g., [
7]). However, it is possible to introduce a norm on a given ring that has natural properties and induces a metrizable topology. Let us recall the following definition (cf. [
33]).
Definition 3. If R is any ring, then a real-valued function defined on R is called a norm for R if it satisfies the following conditions for all :
- 1.
and if and only if
- 2.
- 3.
- 4.
for some constant
Definition 4. Let us define a norm on in the following way:where is an irreducible representative of Proposition 5. The norm in Definition 4 is well defined on and satisfies the conditions of Definition 3. In addition, Proof. Note that an irreducible representative of is not unique in general. However, if and are irreducible representatives of then they consist of the same coordinates (up to a permutation of nonzero coordinates), and so, Thus, the norm is well-defined.
Clearly, if then is its irreducible representative, and so, Otherwise, The second property of the norm evidently follows from the corresponding triangle property of the norm on a linear space. In addition,
For any representative
of
, we have that
where
is an irreducible representative of
So,
Let
and let
and
be corresponding irreducible representatives. Then, by Proposition 3,
Thus, satisfies Condition 4 in Definition 3 for In addition, by Proposition 3, we can put if or □
We define a metric
on
associated with the norm in the natural way. Let
be in
We set
It is well known and easy to check that is a metric.
Example 1. Let be a sequence in such that as If then if and only if for all values of n that are big enough. Indeed, if then On the other hand, if then as
Proposition 6. The quotient map is discontinuous as a map from the Banach space to the metric space at each point of , except for zero.
Proof. Example 1 can be easily modified to show the discontinuity of the quotient map at any nonzero point. Indeed, let
; then, without loss of generality, we can assume that
Consider
Then,
in
as
but
and so the quotient map is discontinuous at
On the other hand, if a sequence
tends to zero, then
as
, and thus, the quotient map is continuous at zero. □
Theorem 1. The metric space is complete.
Proof. Let
and
be in
and let
be an irreducible representative of
We claim that there exists an irreducible representative
such that in
Indeed, let
be any irreducible representative of
The inequality
implies that there is an irreducible representative
of
such that
Note that
is not necessary irreducible. However, since both
and
are irreducible, it may happen that some coordinates of
y are the same as some coordinates of
d and that some coordinates of
x are the same as some coordinates of
Let us construct
such that
is obtained by permutating the coordinates of
d, and
is obtained by permutating the coordinates of
b, so the coordinates of
d that are equal to some coordinates of
y have the same positions in
as the corresponding coordinates in
y, and the coordinates of
b that are equal to some coordinates of
x have the same positions in
as the corresponding coordinates in
Then,
and
Let
be a Cauchy sequence in
Taking a subsequence, if necessary, we can assume that if
and
then
Let us choose irreducible representatives
of
with
Thus, if
and
then
Hence, is a Cauchy sequence in , so it has a limit Let be the ith coordinate of that is, if and if Clearly, as We claim that if then there is a number N such that for every Indeed, if it is not so, then for every that is big enough, , and we have a contradiction.
For a given
, we denote by
a vector in
such that
has a finite support,
or
, and
Note that for this case,
Let
N be a number such that for every
for all
and
So,
Therefore, is the limit of , and thus, is complete. □
From the triangle and multiplicative triangle inequalities of the norm, we have that the algebraic operations are jointly continuous in
Indeed, let
and
; then,
and
The continuity of the addition implies that if is an additive map from to an additive topological group and is continuous at zero, then it is continuous at any point.
3. Homomorphisms and Supersymmetric Polynomials
Let be a closed multiplicative semigroup of another Banach algebra and let Y be a Banach space with a symmetric basis.
Theorem 2. Let γ be a multiplicative map from to If there is a constant such that then there exists a continuous ring homomorphismdefined by Proof. It is clear that
is additive and does not depend on the representative. In addition,
Let
and let
be its irreducible representative. Then,
Hence, is continuous at zero, and according to the additivity, it is continuous at each point of
By the multiplicativity of
□
Note that in Theorem 2, we do not need the continuity of
Example 2. Let be an open unit ball centered at the origin of a Banach algebra and where is the unity of and is an open ball of radius which is centered at the origin of In addition, let We define by Then, satisfies the conditions of Theorem 2 and, thus, is continuous.
Corollary 1. Any continuous homomorphism φ from a Banach algebra to a Banach algebra can be extended to a continuous homomorphism from to for any infinite-dimensional Banach space Y with a symmetric basis.
Proof. Since
is a continuous linear and multiplicative operator from
to
it follows that
Hence,
satisfies the conditions of Theorem 2 for
; thus,
is a continuous homomorphism from
to
The map
is an embedding of
to
and
Thus, we can consider as an extension of Note that is not a homomorphism of rings because it is not additive. □
The following example shows that for some cases, the condition is not necessary for the continuity of
Example 3. Let for let , and let n be a natural number, We set Then, for every Banach algebra the mapping from to is a continuous homomorphism. Indeed, since for every and Thus, is continuous at zero and, thus, continuous.
Example 4. Let Then, maps to and it is continuous and additive. If the norm is multiplicative, then is multiplicative.
Note that if
is a homomorphism from
to
and for every
for some
then the map
is multiplicative. However, we do not know if every homomorphism from
to
is of the form in Theorem 2.
Let us consider vector-valued mappings on
Let
E be a linear normed space. We say that a mapping
is
supersymmetric if
whenever
In fact, every supersymmetric function can be defined on
by
It is easy to check that if
f is of the form
where
is a map from
to
then
is supersymmetric and additive. If
is multiplicative, then
is so.
Example 5. Let be an irreducible representative of We set Then, f is a supersymmetric complex-valued function.
If
is a Banach algebra, then
is a Banach space, and we can consider
supersymmetric polynomials on
that is, polynomial mappings to a normed space
E that are supersymmetric. Let us recall that a mapping
from a normed space
Z to
E is an
n-homogeneous polynomial if there exists a multi-linear mapping
on the
nth Cartesian degree
of
Z such that
A finite sum of homogeneous polynomials is a polynomial. Continuous polynomials on Banach spaces were studied by many authors (see, e.g., [
31]). The following example gives us supersymmetric polynomials on
for
Example 6. Let for some and For any integer , we define Clearly, polynomials are supersymmetric. Since the mapping is multiplicative and mappings are continuous ring homomorphisms from to
A polynomial
P on
is
separately symmetric if
P is invariant with respect to all permutations
acting by
and
Clearly, if P is supersymmetric, then it is separately symmetric, but the inverse statement is not true.
Example 7. Evidently, P is separately symmetric. Moreover, However, P is not supersymmetric. Indeed, while However, Thus, P has different values on equivalent vectors, and thus, it cannot be supersymmetric.
The minimal algebra generated by polynomials
was studied in [
7,
29] for the case of
and
The next theorem shows that every supersymmetric polynomial can be represented as a finite algebraic combination of polynomials
Theorem 3. Let P be a supersymmetric polynomial on Then, P is an algebraic combination (that is, a linear combination of finite products) of polynomials
Proof. Let
P be a supersymmetric polynomial on
; then,
is separately symmetric. According to [
34],
P is an algebraic combination of polynomials
and
where
Thus, we have
for some constants
Clearly,
Denote
Then, there is a polynomial
such that
According to our assumption,
We can see that
for every
It is known that for every
, there exists a vector
such that
(see, e.g., [
19]). Thus, for every
,
However, this means that q does not depend on Hence, P is an algebraic combination of polynomials □
In particular, in [
29], it was proved that
in
if and only if
for all
The next example shows that in a more general case, supersymmetric polynomials do not separate points of
Example 8. Let and be the algebra with respect to the coordinate-wise multiplication. Then, the vectoris not equivalent to but Let be a complex homomorphism of and let be a ring homomorphism from to ; then, is a ring complex homomorphism of From the following example, we can see that there are complex homomorphisms of constructed in a different way.
Example 9. Consider the case , as in Example 8. For arbitrary , we setwhere Note that Polynomials are of the form (2) for , and the map γ is multiplicative. So, are continuous complex homomorphisms. Polynomials
in Example 9, which are restricted to elements
, are called
block-symmetric polynomials on
(see, e.g., [
4,
23,
26]) or
MacMahon polynomials in the literature [
35].
Example 10. Let and let be the algebra of all square matrices for some fixed Then, is a noncommutative ring of matrix multisets. Let D be the following map from to : Since the determinant ia a multiplicative mapping, D is a homomorphism. The continuity of D follows from the fact that
4. Discussions and Conclusions
We considered the ring of multisets
consisting of elements in a given multiplicative semigroup
of a Banach algebra
and endowed with some natural “supersymmetric” operations of addition and multiplication. We constructed a complete metrizable topology of
generated by a ring norm. In addition, we investigated homomorphisms of
and their relations with supersymmetric polynomials. Note that
is not a linear space over
or
because there is no natural multiplication by scalars (see, e.g., [
7]).
Rings of multisets may have wide applications in neural networks and machine learning. Computer algorithms are often invariant with respect to permutations of input data instances. This observation suggests the use of permutation-invariant sets instead of vectors of a fixed dimension for the organization of input data (see, e.g., [
12]). For this purpose, multisets (sets with possible repetitions of elements) are actually more suitable. However, classical multisets have a poor algebraic structure. For example, a very important operation of the union of two multisets has no inverse. On the other hand, we can consider a set of multisets as a natural domain of symmetric functions (with respect to permutations of variables) that are defined on a linear space. Since the union of multisets does not preserve cardinality, it is convenient to use infinite-dimensional linear spaces of sequences, such as Banach spaces with symmetric bases. All symmetric functions on
X can be extended to the set of multisets, and if
then symmetric polynomials separate different points of the multisets. To get an operation that is inverse to the union, we have to use Grothendieck’s well-known idea, which is widely used in
K-theory. It leads to the construction of classes of equivalences of pairs
where
y plays the role of a “negative part” (while components of both vectors
x and
y are complex numbers or, in the general case, elements of an abstract Banach algebra
). If we consider
x as vector coding information, then
y consists of “negative” information in the sense that if both
x and
y contain the same piece of information (the same coordinate), then this piece of information will be annulated. Therefore, the union can be extended to a commutative group operation on the classes of equivalence, and together with a natural symmetric multiplication, they form a ring structure on the set of classes. Such a ring of multisets of complex numbers was considered in [
7] for the case of
. In this paper, we investigated the situation when the “coordinates” of
x and
y were in a Banach algebra
and sequences of their norms belonged to a Banach space
X with a symmetric basis. It is interesting that the basic results in [
7] can be extended to the general case. In particular, the ring
that was obtained is a complete metric space in a metrizable topology, and it is naturally induced by norms of
and
The main difference is that supersymmetric polynomials separate points of
, while in the general case, they do not.
One can compare the rings of multisets and fuzzy sets. In a fuzzy set, each element may have a partial membership (between 0 and 1) [
36]. In a ring of multisets, elements may have multiple memberships, and even negatively multiple memberships. Note that the ring
is never algebra, even if
(see [
7]). However, it is known [
33] that under some natural conditions, any metric ring
R can be embedded into a normed algebra over the field of fractions over
It would be interesting to construct such an algebra for the ring
and compare it with fuzzy sets and other algebraic structures.