1. Introduction
Throughout this paper, all rings are assumed to be associative. The set is a left ideal of a ring R, called the left annihilator of S in In an analogous way, one can define a right annihilator of S in ring R. A lattice (complete lattice) is a poset P in which every pair (every subset) of P has both a least upper bound and a greatest lower bound in The set of annihilators in a ring forms a lattice whose structure reflects the internal properties of the ring.
In contrast to the lattices of one-sided ideals, lattices of one-sided annihilators possess a very desirable property: symmetry. More precisely, the lattice of left annihilators in a given ring is anti-isomorphic to the lattice of right annihilators. In the case of commutative rings, this implies that the lattice of annihilators is self-dual. Therefore, if we present Hasse diagrams of the lattices of left and right annihilators side by side, the overall structure exhibits central symmetry.
This symmetry, i.e., the anti-isomorphism—between the lattices of one-sided annihilators, allows us to study only the lattice of left (or right) annihilators, with the results being directly applicable to right (respectively left) annihilators as well.
In the literature, the lattices of annihilators of rings with a multiplicative identity element 1 are the primary focus. Among the characterizations of rings with identity using the concept of annihilators, the following can be distinguished:
Characterizations of rings based on the number of elements in the lattice of annihilators, e.g.,
Rings in which the lattice of one-sided annihilators is a two-element lattice, known as domains (see [
1,
2]).
Rings in which the annihilators of distinct elements are different, known as Boolean rings (see e.g., [
3]).
Characterizations of rings in which the lattice of one-sided annihilators coincides (or almost coincides) with the lattice of one-sided ideals, e.g.,
Rings in which the lattices of one-sided annihilators coincide with the lattices of one-sided ideals, known as dual rings (see e.g., [
4])
Artinian rings, in which every left ideal is a left annihilator and every right ideal is a right annihilator, known as quasi-Frobenius rings (see e.g., [
5,
6]).
Rings in which the left annihilator of any single element (or any subset) of a ring is generated, as a left ideal, by an idempotent element, known as left Rickart rings (or left Baer rings, respectively) (see e.g., [
5]).
Rings in which the right annihilator of any subset of the ring is comparable with every other right ideal, known as a right annelidan rings (see e.g., [
7,
8]).
Characterizations of rings in which the lattice of one-sided annihilators has a relatively simple structure, e.g.,
Rings in which the lattice of one-sided annihilators is a chain, known as lineal rings (see e.g., [
9,
10]).
Rings in which the lattice of one-sided (or two-sided) annihilators is a Boolean lattice (see e.g., [
11,
12] ).
The assumption of the existence of a multiplicative identity element 1 in a ring significantly affects even the properties of lattices of one-sided ideals (see [
2,
13,
14,
15]). The lattice of left annihilators in a ring is a subsemilattice of the meet-semilattice of left ideals, so its properties must also depend on the presence of 1 in the ring. It is also well known that many characterizations of rings with 1 obtained using the notion of annihilator cannot be extended to the case of rings without identity. For example, the statement: ”The lattice of one-sided annihilators in a ring with a 1 is a Boolean lattice if and only if the ring is reduced“ is false for rings without identity (see [
11] ). The left (and right) annihilator of the entire ring with identity is the zero ideal, which need not be the case for rings without identity. Therefore, a natural approach to studying the existence of an identity element in a ring is through the investigation of annihilators.
In this paper, we restrict ourselves to the study of a particular class of rings, namely algebras over fields (briefly referred to as algebras). This class is very important because typical examples of rings—such as polynomial rings over fields, matrix rings over fields, and their subalgebras (both with and without a multiplicative identity element)—are algebras over fields. Moreover, many examples of rings in which the lattice of one-sided annihilators has a specific form (e.g., is not modular) are found within this class (see e.g., [
11,
15,
16,
17,
18,
19]).
In
Section 2, we begin by recalling selected definitions and results concerning algebras, which are needed in the later parts of the paper. In some cases, the cited results are modified or even strengthened compared to their ring-theoretic versions known from the literature (cf. [
1,
5,
20,
21]). Examples include Theorems 2 and 3. Such generalizations are possible because they apply to algebras over fields rather than to arbitrary rings.
Section 3 is devoted to the properties of annihilators in algebras. We examine the relationship between the lattice of one-sided annihilators of an algebra
A and the corresponding lattice of its unital extension
focusing primarily on the case where
A is a semiprimary algebra. In
Section 4, we provide equivalent conditions under which a semiprimary algebra possesses an identity element. We also show which natural conditions are insufficient to ensure the existence of an identity.
All information about lattices used here can be found, for example, in [
15,
22].
2. Introduction to Algebras over Fields
An algebra over a field
, or short an algebra is a ring
having at the same time the structure of a linear space over
with the same addition, where for any
and for any
the following condition is satisfied:
By we denote an algebra with the same linear structure as A over , but with inverse multiplication, say ∗. Thus in we have for By the dimension of the algebra A we mean its dimension as a linear space over . The set of all linear combinations of elements is denoted by
The concept of algebras over fields emerged in the algebraic literature in the late 19th century and was fully developed in the first half of the 20th century. Examples of algebras over fields include the quaternions, matrix algebras over fields and polynomial algebras over fields. Nowadays, the theory of algebras over fields is used, among other things, to study and construct error-correcting codes (see e.g., [
23]). For more detailed information, history and examples we refer the reader to [
20]. To make the article self-contained, in this section, we recall the key concepts related to algebras over fields, supplemented with additional information needed in the subsequent parts of the paper.
In a situation where this does not lead to any misunderstanding, we will assume that an algebra has a given property if it has this property as a ring. We will then use concepts and notations from ring theory. In this sense we understand algebras with an identity. If the algebra A has an identity, we will write it with the symbol , or in short if it is known which algebra we mean. The left ideal of an algebra A is the left ideal of the ring A which is also a subspace in It may happen that for a given algebra A there exist ideals (and therefore one-sided ideals) of A as a ring which are not ideals of A as an algebra.
Example 1. Let be an infinite field and A a nonzero -algebra with zero multiplication. Let be an arbitrary element in If the characteristic is equal to zero, then for any let
If has the characteristic then, since is an infinite field, there is an infinite sequence of elements linearly independent over a simple subfield In this case, for any let
In both cases the sets are ideals of the ring A which are not even one-sided ideals of the algebra The one-sided ideals of the algebra A are only -subspaces in Therefore, if then the only one-sided ideals in the algebra A are 0 and
If I is the left ideal (right ideal, two-sided ideal) of the algebra , then we will write A (, ). The set of all left (right, two-sided) ideals of the algebra A will be denoted (, ).
Below we have a well-known fact (see e.g., [
14,
20,
24]).
Proposition 1. The sets , and , ordered by the inclusion relation are modular and complete lattices. They are sublattices in the lattice of the subspaces Sub.
By standard convention, if and are subspaces of a linear space V and , then the sum of these subspaces is called a direct sum and denoted by .
Now let A and B be algebras. Let us take the linear space It is naturally the direct sum of the linear subspaces and so Let us extend the multiplications in A and in B to the multiplication in assuming for any Then we get an algebra called direct sum of algebras A and which we still denote by In this situation, A and B are ideals in the algebra
If the algebra A can be written as a direct sum of its ideals I and then we get a decomposition of A into a direct sum of algebras in the sense of the previous definition.
Let
be nonempty subsets of
A and at least one of them is not single-element. Then we denote by
the subspace in
A generated by all products
where
If at least one of these sets is a subspace in
then it is clear that the subspace
is given by the formula
If B and C are subalgebras in A, then need not be a subalgebra. However, if B is a left ideal (C is a right ideal) of A, then their product is also a left ideal (right ideal) in So if B and C are ideals in then obviously is also an ideal in and
Since multiplication on A is associative, the multiplication of subsets of A defined above is also associative. Therefore, this product can be extended to any number of factors, without specifying the parentheses. We assume that The product of factors, each of which is equal to B, will be written as .
A subset B of A is called nilpotent if for some . If moreover and then we say that B has degree of nilpotence It is known that if is a one-sided nilpotent ideal, then the ideal generated by I in A is also nilpotent, of the same degree of nilpotence. Algebras without nonzero nilpotent ideals are called semiprime.
We will now distinguish certain elements of algebras that have important properties. By
center of algebra A we will understand the set
The element will be called central element. If all elements in A are central, then A is a commutative algebra. It is obvious that a ring with identity is a -algebra if and only if we are given a homomorphism of the field into that transforms into
An important type of element in algebras is an idempotent. An element is called an idempotent element (or, more briefly, idempotent ) if it satisfies the condition . Obviously, 0 and 1 are idempotents in A. An idempotent e of A is called a left (right) identity if for any a of Of course, the identity is always a left and right identity. On the other hand, if an algebra has a left identity and a right identity, they are the same and are the identity of the algebra.
In the literature, rings in which the set of left ideals with order determined by inclusion satisfies the Descending Chain Condition (DCC) are known as left Artinian rings. Right Artinian rings are defined analogously. Of course, every left (right) Artinian algebra as a ring is left (right) Artinian as an algebra. From the Example 1 it follows that the converse implication need not be true, even when the algebra has dimension 1 and only two ideals. The situation is simple if the algebra A has an identity because then naturally as a subalgebra. Then, all one-sided ring ideals in A are one-sided ideals of A as an algebra, of course from the same side.
The set of all elements in
A that are simultaneously left- and right-invertible will be denoted by
The right ideal
J of the algebra
A is called maximal ideal, if it is a coatom in the lattice
The right ideal
is called regular, if there exists an element
such that
for all
One of the most important ideals in the algebra
A is the ideal that is the intersection of all right maximals ideals, which are also regular ideals. This ideal is known as Jacobson’s radical or for short radical of algebra
A and is usually denoted by the symbol
Various characterizations of the radical can be found e.g., in [
21]. For us, the following information will be important: for any algebra
its radical is an ideal of
A as an algebra, contains all one-sided nilpotent ideals, and
for any ideal
Moreover, if
A is an algebra with identity, then for any ideal
we have
if and only if
The following theorem is crucial in the theory of noncommutative rings (see e.g., [
1]).
Theorem 1 (Wedderburn–Artin Theorem)
. If R is a semiprime left (or right) Artinian ring, thenwhere each is a division ring and denotes the ring of matrices over Algebras that satisfy the conditions of the above theorem are called semisimple algebras. It should be noted that a semisimple algebra A has an identity element and every one-sided ideal in A is generated by an idempotent element.
Extending the notion of Artinian algebras, an algebra A is called semiprimary if its radical is nilpotent and the quotient algebra is a semisimple algebra. Among other things, all finite-dimensional algebras that are not nilpotent are semiprimary algebras.
Theorem 2. Let A be a semiprimary algebra with radical
- (1)
If is not nilpotent, but for some then there exists an idempotent such that
- (2)
If and then there exists an idempotent such that i.e., the idempotent can be raised to the idempotent
- (3)
If is a one-sided ideal, then it is either nilpotent or contains a nontrivial idempotent element.
Proof. (1) Let
be non-nilpotent, but
for some
Then, we can rewrite this condition in the form
for some polynomial
Hence, by induction, for any
we have
In particular,
Let us put
Then,
so
e is idempotent.
Since a is not nilpotent and then It is also clear from the construction that
(2) If
then it is enough to assume
Let
be such an element that
but
Therefore, it is easy to see that
a is not a nilpotent element. There also exists
such that
If
then it is enough to assume
Let
So since
a is not a nilpotent element, let us take f
and
e as in the proof of point (1). So we have
and
Let ≡ denote congruence modulo
J in
Since
J is an ideal, it follows by assumption that
for any
Therefore,
i.e.,
(3) Let be a one-sided ideal. If then I is nilpotent, since A is a semiprimary algebra. If, on the other hand, then in the semisimple algebra we have a nonzero one-sided ideal By Theorem 1 it contains a nonzero idempotent element, the image of some element Now we see that a is not a nilpotent element, but is a nilpotent element. The thesis follows from point (1). □
We now present a theorem, known in a slightly weaker form for rings, which allows us to take advantage of the properties of algebras with identity.
Theorem 3. Let A be a -algebra. On the direct sum of linear spaces let us determine the multiplication by the formula:where and . Then, is an algebra with identity Using natural identification of A and with the corresponding subspaces in the following conditions are satisfied: - (1)
A is an ideal in and
- (2)
Every left (right) ideal in A is a left (right) ideal in as a ring;
- (3)
If I is a one-sided ideal in the algebra and then there exists an element such that
- (4)
If are left (right) ideals in such that and then
- (5)
A is a left (right) Artinian algebra if and only if is a left (right) Artinian ring.
Proof. It is easy to check that the multiplication given in the Formula (
2) transforms
into a
-algebra with identity equal to
.
The proofs of points (1) and (2) follow from direct calculations. As an example, we compute that every left ideal I in A is a left ideal in as a ring. Obviously, is closed under addition. It follows from Formula (2) applied to that Hence, I is a left ideal in
(3) Let Since I is a subspace in we can assume for some We will show that Let us take any We can write it in the form for some and Of course and which completes the proof.
(4) From the construction of , it follows that A is both a left and a right maximal ideal in Therefore, by assumption, it follows that Hence, since the lattices of one-sided ideals in are modular according to Proposition 1.
(5) Since is an algebra with identity, then is Artinian as a ring if and only if it is Artinian as an algebra. Now, point (5) easily follows from (4). □
Remark 1. One can also obtain (4) as a direct consequence of (3).
3. Lattices of Annihilators of and
We begin this section by recalling the definitions of annihilators and lattices of annihilators known from the literature, and we present the most important properties of annihilators. We provide an example of an algebra whose annihilators possess significant properties. The chapter concludes with a proof of a new result stating that in a semiprimary algebra with right identity the lattice of left annihilators in A is a sublattice in the lattice of left annihilators in
Definition 1. Let A be an algebra and a nonempty subset.
The left annihilator of S in A is the set:Similarly, the right annihilator of S in is the set: We now give some properties of annihilators, which follow directly from the definition.
Lemma 1. Let A be an algebra and let be subsets of Let I be an arbitrary set and let for any be a subset of the algebra A. Then:
- (1)
If , then and
- (2)
and
- (3)
If , then
- (4)
If , then
- (5)
i
- (6)
If are subspaces of A, then and
- (7)
If , then i
In the following, we will be interested in the set of all left annihilators in the algebra A, which we will denote by the symbol We will also consider the set of all right annihilators in the algebra A.
Considering the associative property of multiplication in A, we note that every left annihilator in A is a left ideal in A and similarly, every right annihilator in A is a right ideal in A.
By the Formula (
1), every left annihilator and every right annihilator in
A is a subspace over
, and thus a one-sided ideal of
This implies that
and
.
Let and be left annihilators in the algebra Then, by the Lemma 1, it follows that the set is a left ideal and a left annihilator in A. On the other hand, the set is a left ideal, but need not be a left annihilator. However, the sets and have a lattice structure.
The set
with order given by inclusion is a complete lattice with operations given for any family of left annihilators
by the condition:
Analogously, the set
with order given by inclusion is a complete lattice with operations given for any family of right annihilators
by the condition:
It is well known that in any ring (and algebra) there is a Galois correspondence of the lattices of right annihilators and left annihilators. The lattice
is anti-isomorphic to the lattice
under the mapping
for
The inverse of
l is the mapping
for
This often allows us to restrict our considerations to the lattice of left annihilators of a given algebra, and to interpret the results in terms of the lattice of right annihilators of this algebra. It is also known that there is no nontrivial identity satisfied by lattices of one-sided annihilators in rings (even if ring has an identity element). In particular, every finite lattice is a sublattice of the lattice of annihilators of a certain ring with an identity element (see [
16], for commutative rings see [
18]).
Proposition 2. Let A be a semiprimary algebra. Then, in the lattice there exists a proper left annihilator I such that Similarly, in the lattice there exists a proper right annihilator J such that
Proof. Since A is semiprimary, then by Theorem 2 (3) in A there exists a nontrivial idempotent By point (3) of Theorem 3 we have and □
We now give an example of an algebra A and a left annihilator such that
Example 2. Let A be an algebra of dimension 2 with basis Let multiplication in the basis be of the form: so the table of the multiplication is Then . Hence is a chain with two elements: Then because A is a maximal ideal in It is obvious that therefore Let us additionally note that and . Using (2) of Lemma 1 we have checked that no left annihilator in A is a left annihilator in
Theorem 4. Let A be a semiprimary algebra. Then, if A has a right identity, then the lattice is a sublattice of the lattice Similarly, if A has a left identity, then the lattice is a sublattice of the lattice
Proof. If and is a right identity of an algebra then by Theorem 3 (3) we have It is easy to check that if are subspaces in such that then at least one of the subspaces is contained in Therefore Additionally, using Lemma 1 (1), we have Obviously, and hence
Now let Of course, is a lower bound of in the lattice . Substituting into the above considerations, we obtain From this it is already clear that the lattice is a sublattice of the lattice
The proof of the statement for right annihilators proceeds analogously. □
4. Some Conditions Suspected of Implying the Existence of an Identity Element in an Algebra
It is obvious that if the algebra A has identity, then the following conditions are satisfied:
- (i)
A has a nonzero idempotent;
- (ii)
- (iii)
i.e., A has zero annihilators on both sides.
We now discuss which of the above conditions, possibly under additional assumptions, ensure that A has an identity element.
The following fact follows directly from the definition of one-sided identities and one-sided annihilators.
Proposition 3. If A has a right identity, then Similarly, if A has a left identity, then
The example below shows that conditions (i) and (ii) do not imply the existence of an identity element in the algebra.
Example 3. Let A be an algebra with a basis as linear space and with the multiplication given by and (see Example 2). Then A does not have an identity element, although it has the left identity. Furthermore,
Now let us check how the condition affects the existence of identity element in
For semiprimary algebras we have the following result (cf. [
21] (Theorem 1.4.3.)).
Theorem 5. Let A be a semiprimary algebra and J its radical. The following conditions are equivalent:
- (1)
Algebra A has a left identity;
- (2)
Every homomorphic image of A has a zero right annihilator;
- (3)
For any the algebra has a zero right annihilator.
Proof. Let e be a left identity in and let be an arbitrary ideal. Then, of course, is a left identity in Therefore, by Proposition 3, we have
Obviously,
Let
for some
By assumption it follows that
and the algebra
has identity. Let
be an element such that
is identity in
By Theorem 2 (2) we can assume that
e is an idempotent. Assume that for
we already know that
is a left identity in
So
Furthermore, by the choice of
e we know that
So
The assumption of the value zero of the right annihilator gives that Therefore for we have , so is the left identity of Now, substituting we get that is the left identity modulo so e is the left identity in □
Corollary 1. Let be an infinite field. If the nontrivial -algebra A is left-Artinian as a ring, then A has a left identity.
Proof. Suppose that Then, every subspace (and even a subgroup of the additive group) in r is a left ideal in From the Example 1 and the left Artinianity of A it follows that this is impossible. Therefore, it must be that A similar observation holds for any homomorphic image of the algebra Therefore, from the theorem above it follows that A has a left identity. □
Theorem 6. Let A be a semiprimary algebra and J its radical. The following conditions are equivalent:
- (1)
Algebra A has an identity;
- (2)
Every homomorphic image of A has a zero right and left annihilators;
- (3)
For any the algebra has a zero right and left annihilators.
Proof. If A has a left identity e and a right identity then and this is an identity of the algebra Now the result is a direct consequence of Theorem 5 and its analogous version for algebras with a right identity. □
Corollary 2. Let A be a semiprimary algebra and J its radical. If and then A has identity.
Proof. Note that as a semisimple algebra has an identity, so both of its annihilators are zero. Since the thesis follows directly from Theorem 6. □
It turns out that already in algebras of dimension the condition does not guarantee the existence of even a one-sided identity in A.
Example 4. Let A be an algebra of dimension 4 over the field whose basis is the elements . Let the multiplication table of the elements of the basis have the following form:· | x | y | z | t |
x | 0 | 0 | 0 | x |
y | z | 0 | 0 | 0 |
z | 0 | 0 | 0 | z |
t | 0 | y | z | t |
It is easy to check that multiplication in A is associative. Directly from the table of multiplication we can see that every element of A is of the form where . Moreover, t is idempotent in Furthermore,
We first show that If for some then From the table of multiplication, it follows that Linear independence of elements gives Then, Now we show that Let for some Then and linear independence of elements gives This shows that Form Lemma 1 (1) and follows and Therefore, as a subspace in satisfies either or Analogously, or
- (a)
and then
- (b)
and then
Therefore, A has zero annihilators on both sides.
Since in any algebra the left (right) identity is an idempotent, let us determine the idempotents in Let be an idempotent in Hence, Then, and we obtain system of equations Hence or If then If then for any
In consequence, is the set of all nonzero idempotents in However, and That means that there is no idempotent which is left identity in A, and analogously there is no idempotent which is right identity in A.
The results obtained appear to provide a promising starting point for further investigation of lattices of one-sided annihilators in rings without identity. It is possible that simple characterizations of such rings may be found in terms of their lattices of annihilators.