The covering numbers of rings
Abstract.
A cover of an associative (not necessarily commutative nor unital) ring is a collection of proper subrings of whose set-theoretic union equals . If such a cover exists, then the covering number of is the cardinality of a minimal cover, and a ring is called -elementary if for every nonzero two-sided ideal of . If is a ring with unity, then we define the unital covering number to be the size of a minimal cover of by subrings that contain (if such a cover exists), and is -elementary if for every nonzero two-sided ideal of . In this paper, we classify all -elementary unital rings and determine their covering numbers. Building on this classification, we are further able to classify all -elementary rings and prove for every -elementary ring . We also prove that, if is a ring without unity with a finite cover, then there exists a unital ring such that , which in turn provides a complete list of all integers that are the covering number of a ring. Moreover, if
then we show that , which proves that almost all integers are not covering numbers of a ring.
1. Introduction
In this article, rings are assumed to be associative, but need not be commutative nor have a multiplicative identity. Our goal is to study the ways in which a ring can be expressed as a union of proper subrings. Here, we use subring in the weakest sense: is a subring when is an additive group and is closed under multiplication. In particular, need not contain a multiplicative identity, even if itself is unital. A cover of is a collection of proper subrings of such that . When such a cover exists, we say that is coverable, and define the covering number to be the cardinality of a minimal cover. If is not coverable, we set .
The origin of this problem comes from group theory, where there is an analogous notion of a cover (by subgroups) for a noncyclic group, and covering problems have a long and fascinating history. That no group is the union of exactly two proper subgroups is an easy exercise (and, in fact, a Putnam Competition problem from 1969; see [13, Chapter 3, Exercise 15]). The earliest known consideration of covering numbers of groups was perhaps by Scorza [30], who proved that a group is the union of three proper subgroups if and only if it has a quotient isomorphic to the Klein -group; a nice proof of this result, which has been rediscovered a number of times through the years, can be found in [1].
Problems related to covering numbers of groups have received considerable attention in recent years. Cohn [8] proved that there exists a group with covering number for every prime and positive integer and conjectured that every solvable group has a covering number of this form. This conjecture was proven by Tomkinson [34], who also proved that there is no group with covering number equal to 7. Covering numbers have been determined for various families of simple and almost simple groups (see, for instance, [2, 3, 4, 17, 24, 31]). As noted above, no group has covering number 7, and other natural numbers exist that are not the covering number of a group (the next smallest examples being 11 [10] and 19 [14]). All the integers that are not the covering number of a group were determined in [15], which also contains a good summary of the history and recent developments of the related work for groups. It is not known whether there are infinitely many positive integers that are not the covering number of a group.
A natural question then arises: what can be said about covering numbers of other algebraic structures? The covering number of a vector space by proper subspaces was determined by Khare [19] (see also [6]). The 2014 article by Kappe [18] provides a survey of recent research, but there have been some notable results since its publication. Gagola III and Kappe [12] proved that for every integer , there exists a loop whose covering number is exactly , and Donoven and Kappe [11] proved that the covering number of a finite semigroup that is not a group and not generated by a single element is always two. Furthermore, they showed that for each , there exists an inverse semigroup whose covering number (by inverse subsemigroups) is exactly . Recently, the covering number of modules by proper submodules was studied in [20] and [16].
The related question for rings has also received significant attention, especially in the last decade. Lucchini and Maróti [23] determined all rings with covering number equal to three. However, as noted by Kappe [18, p. 86], “the solution is less simple than the group case,” and a ring has covering number three if and only if it has a homomorphic image isomorphic to one of five rings. In light of this result, Kappe [18, p. 87] further notes that a characterization of rings which are the union of exactly four proper subrings “does not seem to be a feasible problem for rings.” The difficulty of this problem is further illuminated in recent work by Cohen [7], in which a strategy is presented to classify unital rings with a given covering number and a partial classification is given of unital rings whose covering number is four. Moreover, for each integer , , examples are known of finite rings (with unity) that have covering number , and no integers have thus far been ruled out from being the covering number of a ring. Indeed, Kappe [18, p. 87] further writes, “An interesting question would be if there are integers that are not the covering number of a ring.” Other recent works on this problem include [5, 9, 27, 32, 35].
The goal of this paper is to provide a general method to calculate for any ring with a finite covering number, and consequently to determine all integers that occur as the covering number of a ring. Our main result (Theorem 1.3) allows one to describe the covering numbers of all unital rings. A variation on this theorem (Theorem 1.5, Theorem 1.6) can be used to compute the covering number of any nonunital ring. Thus, we are able to characterize all the positive integers that are the covering number of a ring, and prove that the set of such integers has density 0 (Corollary 1.9).
Our approach to these problems is based on the following easy observation: if has a two-sided ideal such that admits a cover , then we may form a cover of by taking the inverse images of the subrings in under the natural homomorphism . Thus, . Of capital importance are those rings for which the inequality is strict for all .
Definition 1.1.
A ring is said to be -elementary if for every nonzero two-sided ideal of . Note that a -elementary ring must be coverable, since .
Observe that if is finite, then either is -elementary, or has a proper, -elementary residue ring with the same covering number. Moreover, if an integer occurs as the covering number of a ring, then there exists a -elementary ring such that . Thus, many questions about covering numbers of rings can be reduced to the case of -elementary rings.
In a recent paper [33], the authors introduced a new family of -elementary rings, called rings of AGL-type (see Definition 1.2) and determined their covering numbers. In Theorem 1.3 below, we will prove that rings of AGL-type, along with three other infinite families of rings (which have been studied previously), provide a complete list of all -elementary rings with unity. Furthermore, we give formulas for the covering numbers of each family of -elementary rings.
In order to state Theorem 1.3, we must introduce some notation. For a prime power , is the finite field of order , and is the matrix ring with entries from . The Jacobson radical of a ring is denoted by , and when contains unity, the unit group of is . When is a commutative ring and is an -module, the idealization of and , denoted , is the ring
with multiplication . Of particular importance is the idealization of with the 2-dimensional vector space . This ring can be represented as the following ring of matrices:
Next, we give some definitions from [33] related to rings of AGL-type.
Definition 1.2.
Given two powers and of a prime , we define to be the order of the field compositum of and , which is . Observe that when and , we have .
Let and let . We define to be the following subring of :
and call this a ring of AGL-type. The construction and terminology for were inspired by representations of the affine general linear group , which is isomorphic to a subgroup of .
The formulas to compute require some accessory functions. For a prime power , we take to be the set of all monic irreducible polynomials in of degree . We define
We let denote the prime omega function, which counts the number of distinct prime divisors of a natural number. Note that . For , we define
which counts the number of maximal subrings of the field . The variation in the definitions between and is due to the fact that when , is the only maximal subring of ; but, when , the maximal subrings of are exactly the maximal subfields of prime index. Finally, we let denote the -binomial coefficient, which counts the number of -dimensional subspaces of :
We can now state our first main result.
Theorem 1.3.
Let be a -elementary ring with unity. Then, one of the following holds.
-
(1)
If is commutative and semisimple, then for some prime power , and
-
(2)
If is commutative but not semisimple, then for some prime power , and
-
(3)
If is noncommutative and semisimple, then for some prime power and integer , and
where is the smallest prime divisor of .
-
(4)
If is noncommutative and not semisimple, then and one of the two cases below holds. Let , and if , then let be the smallest prime divisor of .
-
(i)
and or . In this case, .
-
(ii)
, , and . In this case,
-
(i)
Portions of Theorem 1.3 have been proved in earlier papers. The classification of commutative unital -elementary rings was done in [32], and the covering numbers of the rings in parts (1) and (2) were determined in [35]. The formula for the covering number of is due to Crestani, Lucchini, and Maróti [9, 23]. Since is simple, it is clearly -elementary, and the fact that these are the only noncommutative semisimple -elementary rings follows from the classification of covering numbers for finite semisimple rings done in [27]. The determination of when is -elementary, and, in that event, the computation of its covering number, were completed in [33]. The content of the present paper is a proof that any -elementary unital ring falls into one of the four classes listed in Theorem 1.3.
While Theorem 1.3 handles the situation where contains unity, it leaves open the problem of determining which integers are covering numbers of nonunital rings. This more general case can be reduced to the unital case by considering covers of in which every subring contains .
Definition 1.4.
Let be a ring with unity. If can be covered by unital subrings (i.e. those containing ), then the unital covering number is defined to be the cardinality of a minimal cover by unital subrings. We say is -elementary if for every nonzero two-sided ideal of . Note that a -elementary ring necessarily contains unity, and admits a cover by unital subrings.
Theorem 1.5.
Let be a ring without unity that has a finite cover. Then, there exists a ring with unity such that . Thus, every covering number of a nonunital ring by subrings occurs as the covering number of a unital ring by unital subrings.
Via Theorem 1.5, we can describe the covering numbers of all rings (with or without unity) by classifying -elementary rings. Our classification is very similar to that of unital -elementary rings, which was given in Theorem 1.3.
Theorem 1.6.
Let be a -elementary ring.
- (1)
-
(2)
If is not -elementary, then either with , or with .
In particular, if is -elementary, then , and the integers that are covering numbers of rings with unity are the same as the integers that are unital covering numbers of rings with unity.
We remark that Theorem 1.6 is likely of independent interest: for many, the definition of a ring includes the existence of a multiplicative identity, and subrings likewise are required to contain this unital element; see, e.g., [21, Chapter II]. (Those who adopt this convention often refer to a ring without unity as an rng.) In this setting, the only covers would indeed be covers by unital subrings.
In light of these results, any integer that is the covering number of a ring (with or without unity) or unital covering number of a ring with unity can be computed by the formulas in Theorem 1.3.
Corollary 1.7.
Any integer that is the covering number or unital covering number of a ring occurs as the covering number of a ring with unity.
This, in turn, allows us to answer Question 1.2 from [32]:
Corollary 1.8.
There does not exist an associative ring with covering number .
Even more, Theorems 1.3, 1.5, and 1.6 allow us to prove that there are infinitely many integers that are not covering numbers (or unital covering numbers) of rings. In fact, almost all integers are not covering numbers of rings.
Corollary 1.9.
Let Then, for all ,
where denotes the binary (base ) logarithm of . In particular, we have and
This paper is organized as follows. For the majority of the article, we will focus on rings with unity. In Section 2, we collect a number of background results that will be used in the rest of the article. Section 3 lays the groundwork for the proof of Theorem 1.3 by studying the Peirce decomposition of the Jacobson radical of a -elementary ring and providing numerical bounds of covering numbers in certain cases. Section 4 is dedicated to proving that most terms in the Peirce decomposition of the Jacobson radical of a -elementary ring must in fact be (so, a -elementary ring cannot be “too far” from being semisimple), and Section 5 in turn places severe restrictions on the direct summands of the semisimple residue of a -elementary ring. This portion of the work culminates in Section 6, where we use the results of the previous sections to prove Theorem 1.3.
In Section 7, we deal with the general case where rings need not contain a multiplicative identity. After proving Theorem 1.5, we concentrate on proving Theorem 1.6, which is accomplished using the methods developed in earlier sections. Finally, in a short post-script (Section 8), we employ the formulas for in Theorem 1.3 to prove Corollary 1.9.
2. Frequently used results
In this section, we summarize a number of results that will be referenced frequently later in the paper. The first lemma is elementary, and will be taken for granted throughout the article.
Lemma 2.1.
Let be a ring with unity.
-
(1)
is coverable if and only if cannot be generated (as a ring) by a single element.
-
(2)
If is noncommutative, then is coverable.
-
(3)
For any two-sided ideal of , a cover of can be lifted to a cover of . Hence, .
-
(4)
If each proper subring of is contained in a maximal subring, then we may assume that any minimal cover of consists of maximal subrings.
By Lemma 2.1(4), maximal subrings of are important in forming minimal covers. The next lemma provides conditions under which a maximal subring is guaranteed to be part of every cover of a ring.
Lemma 2.2.
As shown in [35, Theorem 2.2], when is a direct sum of rings, can be determined from the covering numbers of the summands of as long as all maximal subrings of respect the direct sum decomposition. While not explicitly stated in [35], the analogous result for unital maximal subrings and unital coverings of is also true, and can be proved in the same manner as [35, Theorem 2.2].
Lemma 2.3.
[35, Theorem 2.2] Let for rings . Assume that each maximal subring of has the form
for some . Then, . Furthermore, if each is a unital ring and each unital maximal subring has the above form (with itself unital), then .
From results of Neumann [26, Lemma 4.1, 4.4] and Lewin [22, Lemma 1], it is known that if has a finite covering number, then there exists a two-sided ideal of of finite index such that . Hence, the calculation of covering numbers reduces to the case of finite rings. In fact, a much stronger reduction is possible. Recall that denotes the Jacobson radical of a ring .
Lemma 2.4.
[32, Theorem 3.12] Let be a (unital, associative) ring such that is finite. Then, there exists a two-sided ideal of such that is finite; has characteristic ; ; and .
Thus, for unital rings with finite covers, the computation of reduces to the case where is finite of characteristic and . When is finite, is semisimple, i.e., it is isomorphic to a direct sum of matrix rings over finite fields. The Wedderburn-Malcev Theorem [28, Sec. 11.6, Cor. p. 211], [25, Thm. VIII.28] (sometimes called the Wedderburn Principal Theorem) provides a more detailed description of the structure of , and how it relates to .
Theorem 2.5.
(Wedderburn-Malcev Theorem) Let be a finite ring with unity of characteristic . Then, there exists an -subalgebra of such that , and as -algebras. Moreover, is unique up to conjugation by elements of .
Let be the set of all semisimple complements to in . By Theorem 2.5, all such complements are conjugate. So, for any , we have , and .
The decomposition given by Theorem 2.5 allows us to characterize all of the maximal subrings of .
Lemma 2.6.
[32, Theorem 3.10] Let be a finite ring with unity of characteristic , let be a maximal subring of , and let .
-
(1)
if and only if is the inverse image of a maximal subring of .
-
(2)
if and only if , where and is an ideal of that is maximal among the subideals of contained in .
3. Peirce decomposition and bounds
We now turn our attention to proving that any -elementary unital ring falls into one of the four classes given in Theorem 1.3. Throughout this section, we will assume that is a -elementary unital ring with finite. By Lemma 2.4, we can assume that is finite of characteristic and has Jacobson radical with . Moreover, using Theorem 2.5, we have , where is a semisimple complement to in . The set of all such complements is denoted by . For the remainder of this section, we fix . Then, for some , , where each is a simple ring. For each , let , where and is a power of . Finally, let , so that are orthogonal idempotents such that is the unity of .
Apply a two-sided Peirce decomposition to using the idempotents . This gives , where each is an (, )-bimodule. If and , then (, )-bimodules are equivalent to modules over , where is the compositum of and . In particular, we may speak of the length of as an (, )-bimodule, i.e., the number of simple (, )-bimodules in a direct sum decomposition of . Note that any simple (, )-bimodule has order .
Definition 3.1.
For all , let . If , then let be the length of as an (, )-bimodule. If , then we take .
Recall that when and are powers of , we define to be the order of . If and , then . Note also that we have for any positive integers and . Thus, .
We remark that the rings in parts (2) and (4) of Theorem 1.3 may occur as subrings of , and can be described using the notation and parameters we have defined. If and for some , then . If with , , and , then is a ring of AGL-type.
As we now show, any (, )-bimodule inside of is a two-sided ideal of , and has an ideal complement in .
Lemma 3.2.
Let .
-
(1)
Let be an (, )-bimodule. Then, , is a two-sided ideal of , and has a two-sided ideal complement in .
-
(2)
Let be a two-sided ideal of . Then, has a two-sided ideal complement in .
-
(3)
Let be a two-sided ideal of and let . For all , if and only if . Hence, the number of conjugates of in is equal to .
-
(4)
For each , has a residue ring isomorphic to . For all with , has a residue ring isomorphic to .
Proof.
(1) Since and are the identities of and , respectively, . To show that is a two-sided ideal of , let and . Then, for some and . Since , . Next, write , where each . Since the idempotents are orthogonal, we have . So, . The proof that is similar. Finally, since and are simple rings, is a semisimple (, )-bimodule. So, there is an (, )-bimodule such that , and by the work above, is also a two-sided ideal of .
(2) Apply the Peirce decomposition to to get . Then, each is an (, )-bimodule contained in . By part (1), each has an ideal complement, and therefore does as well.
(3) Note that since , for all . So, for all and all . Certainly, if , then . Conversely, assume that is such that . By (2), there is an ideal complement of in . Express as , where and . Then, for all ,
and because . It follows that . Thus, is central in and .
(4) We will prove the claim regarding . The proof of the other statement is similar. Let , and let , which is a direct sum complement to in . Let be an ideal complement to in , which exists by part (2). Since the idempotents are orthogonal, and annihilate one another. Hence, is a two-sided ideal of , and . ∎
Proposition 3.3.
Assume that is -elementary. Let be a maximal subring of that does not contain . Then, is contained in every minimal cover of .
Proof.
Proposition 3.3 (often in conjunction with Lemma 2.2) will be used to produce lower bounds on the covering number of a -elementary ring. The remainder of this section contains various bounds on the covering numbers of and its subrings.
Lemma 3.4.
Let with , and let .
-
(1)
.
-
(2)
Let be a subideal of . Then, for all , if and only if .
-
(3)
.
Proof.
The next two lemmas are generalizations of [33, Proposition 3.3].
Lemma 3.5.
Let and let . Assume that , and let . Then, .
Proof.
For convenience, let and . First we argue that we may assume is a simple (, )-bimodule. Indeed, if is not simple, then let be a simple (, )-bimodule. By Lemma 3.2, has an ideal complement in . Since , we get . Hence, we may assume that is simple.
For each nonzero , define . Then, each is a subring of . By Lemma 3.4(1), for all nonzero , so is a proper subring of for all .
Let . We will show that is a cover of . Suppose ; that is, suppose for some .
We claim that . We know that for all , so . Furthermore, if and only if . Since for all , this means that if and only if . Hence, , and so . Thus,
and is a cover of .
Finally, note that whenever . This means that there are at most subrings of . Since is a simple -bimodule, . Therefore,
as required. ∎
Lemma 3.6.
Let such that and , and let . If is a simple (, )-bimodule and is -elementary, then . If, in addition, , then and .
Proof.
Assume that is simple and is -elementary. Let . Then, for each , is a maximal subring of . By Lemma 3.4, , so there are distinct conjugates of in .
Let and let be an ideal complement to in . Then, for any conjugate of in , we have
Since is simple, each subring is maximal. On the one hand, because is -elementary, each such subring is contained in every minimal cover of by Lemma 2.2. So, . On the other hand, the union of all the conjugates cannot cover . To see this, let be the intersection of all conjugates of . Then, the union of all the conjugates covers at most
elements of . It follows that the union of all the subrings cannot cover . Hence, .
Recall the definitions for and stated prior to Theorem 1.3. It is shown in [35, Theorem 3.5] that a direct sum of copies of is coverable if and only if .
Lemma 3.7.
Let , let , and let . If , then . If , then .
Proof.
As stated in Theorem 1.3(1), we have
where is equal to 1 if , and is equal to the number of prime divisors of otherwise. Via this formula, it is clear that when . So, assume that . Then, , and
(3.8) |
Now, it is well known that , with equality if and only if . When , , and hence . Combining this with (3.8) produces the desired upper bound for . ∎
4. Restrictions on
Maintain the notation given at the start of Section 3. Our goal in this section is to prove that when is -elementary, most of the bimodules must be zero. In essence, this means that a -elementary ring is “almost” semisimple. We will first examine the bimodules with , and then consider the (, )-bimodules .
Theorem 4.1.
Let with . Assume that and is -elementary. Then, the following hold:
-
(1)
for all such that and .
-
(2)
is a simple (, )-bimodule.
-
(3)
At least one of or is a field.
Proof.
Throughout, let and .
(1) Let be a maximal subideal of . If is a conjugate of in , then is a maximal subring of . Let be an ideal complement to in , and let
which is a direct sum complement to in . Then, for any conjugate of in ,
(4.2) |
is a maximal subring of . The number of such subrings is equal to the number of distinct maximal subrings of of the form . Since the size of a simple -bimodule is , by Lemma 3.4(2), contains maximal subrings of the form in (4.2).
Suppose that for some . The arguments of the previous paragraph can be applied to . Employing the analogous notation and letting , we see that contains maximal subrings of the form
(4.3) |
where is a conjugate of . Moreover, no ring of the form of (4.2) can equal a ring of the form of (4.3), because . By Proposition 3.3,
(4.4) |
(4.5) |
Thus, by (4.4) and (4.5), . This gives
However, we also have . Working as above, this leads to , a contradiction.
(2) Suppose that . Then, still contains maximal subrings as in (4.2). But, contains at least two distinct maximal subideals, so we have at least two choices for . Thus, contains at least maximal subrings of the form of (4.2). Appealing again to Lemma 3.5, we have
But then, , which is impossible. Hence, and is simple.
(3) Suppose that both and are noncommutative, i.e., that and . As in part (1), we obtain by considering maximal subrings of the form in (4.2). Without loss of generality, assume that . Since is -elementary, . But,
a contradiction. Thus, at least one of or must be a field.
∎
Remark 4.6.
It remains possible that is -elementary and for some . Indeed, this case cannot be excluded entirely, because is -elementary and has a nonzero radical. However, we can restrict the possibilities for in this case as well. First, we show that if is -elementary and is noncommutative, then .
Lemma 4.7.
Let and let . Assume that and . Then, .
Proof.
For convenience, let , , and . As in [28, Section 11.8], decomposes as a direct sum of simple (, )-bimodules . Each is equivalent to a simple module—hence has order —and the isomorphism type of is determined entirely by . Thus, we can represent each (additively) as a copy of , and represent in block matrix form as
Via this representation, we see that if and only if and each is an scalar matrix. Thus, .∎
Proposition 4.8.
Let and let . Assume that .
-
(1)
is coverable, and .
-
(2)
If is -elementary, then .
Proof.
As in earlier proofs, we will drop some subscripts for readability. Let , , and .
(1) The ring is coverable because it is noncommutative. For the covering number of , we use induction on . If , then and we are done. So, assume that and that the covering number of is whenever is an (, )-bimodule of length .
Consider , where is a maximal -subideal of . Note that has length , so and . On the one hand, has an ideal complement in , so occurs as a residue ring of and . On the other hand, by Lemmas 3.2(3) and 4.7, the number of conjugates of in is
By [33, Lemma 3.4], , so any minimal cover of excludes some conjugate of . Hence, by Lemma 2.2(1).
(2) Suppose that is -elementary, but . Let be a maximal subideal of , let be an ideal complement for in , and let . Then, is maximal in , so
is a maximal subring of . As noted earlier, the number of conjugates of in is . So, contains such maximal subrings . Since
every minimal cover of excludes at least one . By Lemma 2.2(1), . But, occurs as a residue ring of , so , a contradiction. ∎
Next, we consider when is a field. We will prove that for to be -elementary in this case, it is necessary that ; and, if for some , then whenever is isomorphic to or .
Lemma 4.9.
Let . Assume that is a field and that is a simple (, )-bimodule.
-
(1)
contains a unique maximal subring that does not contain .
-
(2)
and is not -elementary.
Proof.
(1) First, we note that . Indeed, suppose that and . Write , where each . Then, . But, is commutative, so .
Next, let be an ideal complement for in , and let . Then, is a maximal subring of because is simple. Let be any maximal subring of that does not contain . By Lemma 2.6, for some and some maximal subideal of .
We claim that . To see this, apply the Peirce decomposition to to get
Since is simple and , . Then, by maximality, .
Now, let and be such that . Since is central, conjugation by fixes elements of pointwise. Thus,
and so .
(2) As in part (1), let be the ideal complement for in and let be the unique maximal subring of that does not contain . Let be a minimal cover of , and let be all the maximal subrings in that contain .
Note that . Let , let , and consider . Since and is simple, if then , a contradiction. So, , and hence as well. Consequently, . This holds for all , so . Recalling that for each , we see that is a cover of . Thus, , and so in fact . ∎
Remark 4.10.
Proposition 4.11.
Let such that . Assume that is a simple (, )-bimodule, is a field, and is -elementary. Then, for all , if , then .
Proof.
Assume is such that , and suppose that . Since is -elementary, cannot be simple by Lemma 4.9. So, .
Based on our work so far, it remains possible that and for some field that is isomorphic to neither nor . Ultimately, this case is ruled out by our classification of -elementary rings in Theorem 1.3. Rather than attack this case directly, however, we merely note that is always a commutative ring. The conclusion that when will be reached later by referencing the classification of commutative -elementary rings from [32].
5. Restrictions on direct summands of
Continue to use the notations of Sections 3 and 4. At this point, we know that in a -elementary ring, the radical has a restricted structure. There is at most one pair with and , and if , then must be a field not isomorphic to or . We can also place restrictions on the direct summands of . In Proposition 5.3, we show that if is noncommutative, then for all . A closely related result holds for a field (Proposition 5.6): if with and is a field, then for all .
We begin by recalling some results from [27].
Lemma 5.1.
Lemma 5.2.
Let . Assume that for all . Then, is a two-sided ideal of .
Proof.
Let and
so that . By assumption, for all . So,
This means that for all and . It follows that is a two-sided ideal of . ∎
Proposition 5.3.
Let such that . If is -elementary, then for all .
Proof.
Assume that is -elementary, but there exists such that . By Theorem 4.1, there is at most one pair with and , and we must have because and are noncommutative. Also, by Proposition 4.8. It follows that either for all , or for all . Assume without loss of generality that the latter condition holds. Then, is a two-sided ideal of by Lemma 5.2.
For any -algebra isomorphism , let be the maximal subring of . Note that if and only if . Let
which is maximal subring of . Furthermore, since , we have
and so .
Now, let be the number of maximal subrings of isomorphic to . Then, is greater than or equal to the number of choices for , which by Lemma 5.1(2) is greater than . Since has a residue ring isomorphic to , we have . Hence, any minimal cover of omits some maximal subring isomorphic to . Applying Lemma 2.2(1), we have , which is a contradiction. ∎
Ruling out fields isomorphic to is more difficult. By [35, Theorem 3.5], is coverable if and only if , where is as defined in the introduction. When is not coverable, it is generated (as a ring) by a single element . By [35, Proposition 3.4], generates if and only if each is a primitive element for over (if , then any nonzero element of is primitive) and each has a distinct minimal polynomial over .
We also require knowledge about the maximal subrings of . These subrings were classified in [27].
Definition 5.4.
[27, Definition 4.3] Let , where each is a finite simple ring of characteristic . Let be a maximal subring of . We say that is of Type if there exists an index and a maximal subring of such that
We say that is of Type if there exist indices and an -algebra isomorphism such that
here, there are no restrictions on the summands for . In a Type maximal subring, we say that and are linked by .
Lemma 5.5.
[27, Theorem 4.5] Any maximal subring of is Type or Type .
Proposition 5.6.
Let with . Assume that and is a field. If is -elementary, then for all .
Proof.
Assume that is -elementary. By Theorem 4.1, is a simple (, )-bimodule. Let . Suppose that , let , and let . From Theorem 4.1, we know that whenever and . Moreover, by Proposition 4.11, for all . Hence, by Lemma 5.2, each —and therefore —is a two-sided ideal of .
Let and . If , then by Lemma 3.6, and there is nothing to prove because . So, we may assume that .
We break the proof into two cases, depending on whether or . The latter case is easier, so we will deal with that first.
When , is coverable, and since occurs as a residue ring of . By Lemma 3.7,
(5.7) |
However, by Lemma 3.6, . This is not compatible with (5.7), since neither nor is a valid inequality.
For the remainder of the proof, assume that . Define to be . Then, .
We will sort the maximal subrings of into three classes , , and . Let be the collection of all maximal subrings of that contain . Note that these maximal subrings are all inverse images of maximal subrings of .
Next, suppose is maximal with . If , then for some and some maximal subideal . However, as noted earlier, for all and all . So, the action of on is trivial. It follows that for all , and hence must be contained in .
Since and , is the inverse image of a maximal subring of that does not contain . In light of Lemma 5.5, we can partition such subrings into two classes. Use a bar to denote passage from to . Let be the set of all subrings of the form , where is maximal in . Subrings in are those where is not linked to any , . Finally, let be the collection of maximal subrings where is linked to some with .
Now, since , is not coverable. Next, we will show that for all , we can find such that generates (as a ring), and is not contained in a maximal subring in either or . This is easy to see with regard to : for any and as above, if , then for some maximal subring of . But, this is impossible because generates .
For , the argument is more delicate. Let be an additive complement to in . Write as , where and . Since , is less than or equal to the number of irreducible polynomials in of degree . So, for each , we can find such that generates ; each has a distinct minimal polynomial over ; and no has the same minimal polynomial as . Let . Then, for every and every automorphism of , . We conclude that cannot lie in any subring where is linked to some .
We have shown that for every , there exists such that lies in a maximal subring in class . Now, consider a minimal cover of . Let . By our work above, , because elements of the form can only lie in maximal subrings in . Furthermore, the subrings in form a cover of , so . This contradicts the fact that is -elementary, and we are done.∎
6. Proof of Theorem 1.3
We now have all the ingredients necessary to prove Theorem 1.3. Let be a -elementary ring with unity. The idea of the proof is to write as a direct sum of subrings in such a way that all maximal subrings of respect the decomposition. That is, any maximal subring of is equal to one of
(6.1) |
where is maximal in . Then, by Lemma 2.3, and the problem is reduced to a small number of possible cases.
Proposition 6.2.
Assume is a finite unital ring of characteristic containing subrings , , and that satisfy all of the following properties:
-
•
.
-
•
for some .
-
•
, where is some finite index set, for each , and for all .
-
•
is commutative, and if a field occurs as a direct summand of , then .
Then, .
Proof.
It suffices to show that every maximal subring of decomposes as in (6.1). Then, Lemma 2.3 can be used to reach the desired conclusion.
Let . For , let , and fix a semisimple complement for in . Note that and because is semisimple. So, and is a semisimple complement to in .
Let be a maximal subring of . Assume first that . By Lemma 2.6, , where is a conjugate of and is a maximal subideal of . Any such decomposes as either or , where (respectively, ) is maximal in (respectively, ). Moreover, the conjugates of have a similar form. Indeed, has a trivial action on , and likewise for the action of on . So, given with and , we have .
With these observations in mind, we see that if , then
for some conjugate of and some conjugate of . In this case, is maximal in , and . Thus, has the form of (6.1). The same steps show that this also holds when .
Now, assume that . Then, for some maximal subring of . By Lemma 5.5, is either of Type or Type . If is of Type , then has one of the forms
where is maximal in for , and decomposes as in (6.1). On the other hand, if is of Type , then (modulo ) two simple summands and of are linked by an -algebra isomorphism. If is a field, then both and are subrings of , because no simple summand of is isomorphic to . In this case, . Similarly, if is noncommutative, then both and are subrings of , because no simple summand of is isomorphic to . Hence, . Thus, in all cases has the form of (6.1), and we are done. ∎
Proof of Theorem 1.3.
We may assume that is finite, has characteristic , and satisfies . We will use the notation established at the beginning of Section 3. By Theorem 4.1, there is at most pair such that , , and . If such a pair exists, then is simple and one of or is a field. Without loss of generality, assume that . Then, . If , then by Lemma 3.6 and we are done. So, assume that .
If is such that is noncommutative, then by Proposition 4.8, and by Proposition 5.3, for all . In particular, and when . Likewise, by Proposition 4.11, and for all such that , we have .
Form the following three subrings of (these subrings may be equal to ). Let . Let and let . Finally, let and let . Then, , and the subrings satisfy the hypotheses of Proposition 6.2. Thus, . But, is -elementary, so in fact for some , and the other two subrings must be . If , then . If , then is noncommutative and semisimple. Hence, for some by [27, Proposition 5.4, Theorem 5.11]. Finally, if , then is commutative. We may then apply [32, Theorem 4.8] to conclude that either , or , depending on whether is semisimple or not. ∎
7. Rings without unity and covers by unital subrings
Theorem 1.3 provides a complete classification of -elementary rings with unity, which in turn allows us to describe all the integers that occur as covering numbers of such rings. We now examine two related covering problems. First, what can be said of covering numbers for rings without a multiplicative identity? Second, how do our results change if we insist that each subring in a cover of a unital ring must contain ? As shown in Theorem 1.5, these two problems are closely related, and an answer to the second question provides an answer to the first.
Recall the definitions of and -elementary given in Definition 1.4. It is clear that parts (3) and (4) of Lemma 2.1 carry over to coverings of by unital subrings: for any two-sided ideal of , we have ; and when is coverable by unital subrings that are contained in maximal subrings, we may assume that each subring in a minimal cover is maximal. Going forward, we will make free and frequent use of these properties.
It is well known that any ring without unity can be embedded in a larger ring that contains a multiplicative identity. Using this construction, we will connect covers of rings without unity to covers of unital rings by unital subrings and prove Theorem 1.5.
Proof of Theorem 1.5.
We first proceed as in [32, Section 3] and reduce to the case where is finite and has prime power order. By results of Neumann [26, Lemma 4.1,4.4] and Lewin [22, Lemma 1], contains a two-sided ideal of finite index such that . Hence, we may assume that is finite. Next, the Chinese Remainder Theorem still applies to finite rings without identity, so , where each is a ring and for some distinct primes . Lemma 2.3 holds for rings without unity, so . Thus, we may assume that for some prime and some .
Now, as in [32, Lemma 3.2, Proposition 3.3], one may prove that is contained in every maximal subring of . It follows that , and hence we may assume that . From here, we will embed in a unital ring that has characteristic .
Let with multiplication given by the rule:
Notice that, for any , , so has a multiplicative identity. We claim that .
Indeed, let be a minimal cover of by subrings. Given , define , which is a unital subring of , and is proper in because . Let . If , then for some , so . Hence, is a cover of by unital subrings, and so
Next, let be a minimal cover of by unital subrings, which we now know must exist. Let be the projection map defined by . The map is not multiplicative, but is a proper subring of for any . To see this, let . Then, for some . Because , both and are in . Thus, contains both and , and hence . Moreover, , because if not, then and , which implies that .
Finally, given , there exists such that . Hence, and is a cover of of size at most . Consequently, , completing the proof. ∎
In light of Theorem 1.5, determining the covering number of any ring (with or without unity) reduces to the case of a unital ring , for which we may consider either covers by subrings (which need not contain ) or covers by unital subrings. These problems can in turn be solved by classifying -elementary and -elementary rings. The classification of -elementary rings has been done in Theorem 1.3. We now proceed to prove Theorem 1.6, which classifies -elementary rings. As we shall see, the class of -elementary rings is largely the same as the class of -elementary rings, with the greatest discrepancy occurring in the commutative semisimple case.
Lemma 7.1.
Let be a ring with unity. Then, , with equality if and only if the subrings in some minimal cover of each contain .
Proof.
Note that every unital subring is a subring (but not conversely). ∎
Lemma 7.2.
-
(1)
For any prime , is -elementary, but is not coverable by unital subrings.
-
(2)
For any prime , is -elementary, but is not -elementary. Furthermore, .
-
(3)
The ring is -elementary, but is not -elementary. Furthermore, .
Proof.
(1) Let . By [35, Theorem 3.5], is coverable, but is not, so is -elementary. As shown in [35, Theorems 5.3], a minimal cover of requires every maximal subring of . Since is the only maximal subring of , the Type (see Definition 5.4) maximal subrings of do not contain . Thus, is not coverable by unital subrings.
(2) Let for , and let . Given with each , we must have for some . Thus, lies in the Type maximal subring of where is linked to . There are such maximal subrings, and each one contains . Thus, ; but, by Lemma 7.1, so . By part (1), is -elementary. However, has as a residue ring, and thus is not -elementary.
(3) Let , which is equal to the ring of upper triangular matrices over . The three unital subrings , , and form a cover of , so . The nonzero, proper residue rings of are isomorphic to and , neither of which has a unital cover. Hence, is -elementary. Finally, since , is not -elementary. ∎
Proposition 7.3.
Let be a -elementary ring with unity. Then, is -elementary if and only if . Moreover, when is -elementary, we have .
Proof.
We show that, with the exception of , all of the -elementary rings listed in Theorem 1.3 admit minimal covers by unital subrings. Indeed, if , then every maximal subring of contains the identity of the ring [35, Theorem 4.8]. For any , a cover of by unital maximal subrings is given in [35, Example 6.1]. For , maximal subrings of were fully classified in [27, Theorem 3.3], and all such subrings contain the identity matrix. Finally, for rings of AGL-type, minimal covers were constructed in [33]. As noted in [33, Remarks 3.7, 4.3], when is -elementary, it admits a minimal cover by unital subrings. ∎
Proposition 7.3 proves the first half of Theorem 1.6. It remains to show that if is -elementary but not -elementary, then either or . Since we are now focusing on a ring with unity, we will employ many of the same assumptions and notations that were used in Sections 3 through 5. In particular, by Lemma 2.4 it will suffice to consider rings of characteristic with 2-nilpotent radicals. For the remainder of this section, we will assume the following:
Notation 7.4.
Let be a unital ring with characteristic , and let be the Jacobson radical of . Assume that . Let be the set of all the semisimple complements to in . Fix , so that , and write for some simple rings , each with unity . For all , define to be the -bimodule obtained via a Peirce decomposition of . Note that . Let be the length of as an -bimodule; if , then we take .
As in Section 4, our strategy is to show that in a -elementary ring, most of the bimodules must be zero. Much of the work done earlier in this paper still holds when stated only for unital subrings or -elementary rings. As noted earlier, Theorem 4.1, and Lemma 4.9 hold mutatis mutandis for unital maximal subrings and -elementary rings (see Remarks 4.6 and 4.10). In particular, we have the following structural results on a -elementary ring .
Lemma 7.5.
Let be as in Notation 7.4.
-
(1)
If is -elementary, then there is at most one pair with and . For this pair , if , then .
-
(2)
For all , if is a field and , then is not -elementary.
Lemma 7.6.
Let be a -elementary ring that is not -elementary.
-
(1)
There exists a two-sided ideal of such that and .
-
(2)
.
-
(3)
If has a residue ring isomorphic to , then and .
Proof.
(1) Since is not -elementary, it projects nontrivially onto a -elementary ring such that . By Proposition 7.3, all -elementary rings are -elementary except for . If , then , a contradiction to being -elementary. Thus, there exists an ideal of such that . Since is semisimple, .
(2) With as in part (1), we have .
Using Lemma 7.6, we can impose another restriction on and assume that has a residue ring isomorphic to , but does not have a residue ring isomorphic to . Thus, we may assume that has the form
(7.7) |
where for all , and for all .
Lemma 7.8.
Let be as in (7.7) and let be a maximal unital subring of .
-
(1)
If is Type , then contains .
-
(2)
If does not contain , then is Type with linked to for some .
Proof.
(1) Assume that is Type . Then, for some ,
where is a maximal subring of . Suppose that . Then, because . But then, , a contradiction. So, , and .
(2) If is not a subring of , then by part (1) must be Type . In this case, the only way that will fail to contain is if is linked to for some . ∎
Proposition 7.9.
Proof.
(1) Suppose first that for some . Then, (the idealization of with the -dimensional vector space ), which projects onto . By Lemma 3.2(4), has a residue ring isomorphic to . Using this and Lemmas 7.1 and 7.6, we get
Since is -elementary, this forces . But, this contradicts the fact that is not -elementary.
So, we must have for all . However, by Lemma 7.5(2), if for any , then is not -elementary. We conclude that for all .
(2) Suppose by way of contradiction that for some and . By Lemma 7.5(1), there is at most one pair with and . Thus, is unique, and is a simple -bimodule.
Assume without loss of generality that . Let . We claim that is a two-sided ideal of . To see this, note that by part (1), for all . Hence, . Let . Then, , and elements from and mutually annihilate one another. Thus, is an ideal of . Moreover, the action of on is trivial, so is contained in for all . We will show that , which contradicts the fact that is -elementary.
Let be the set of all unital maximal subrings of that contain , and let be the set of all unital maximal subrings that do not. Let , and suppose that . Then, by [32, Theorem 3.10], for some and some maximal subideal . But then, , a contradiction. So, each subring in contains . Use a bar to denote passage to . By Lemma 7.8, is Type with linked to for some , and this is true for every .
From here, we will show that for each , there exists such that is in a maximal subring containing . Let , and write as , where and . Note that is equal to an element of . Let be all of the elements of that are not equal to . Considering as an element of for each , we take .
Certainly, lies in some unital maximal subring of . Suppose that . As noted above, is Type with linked to for some . Since the only automorphism of is the identity, this means that . However, this is impossible, because are all distinct elements of . Thus, lies in some subring in .
To complete the proof, let be a minimal cover of by unital subrings, and let . Then, each subring in contains , and their images in cover . Hence, , which contradicts the fact that is -elementary. ∎
Proposition 7.10.
Proof.
By Proposition 7.9, we know that for all and . Decompose as , where
We claim that all unital maximal subrings of respect this decomposition. That is, if is a unital maximal subring of , then either with maximal in , or with maximal in . When , this can be shown just as in the proof of Proposition 6.2. When , for some maximal subring of . In this case, the only way that would not respect the direct sum decomposition of is if were Type with linked to for some and . However, this is impossible, because for all .
Since every maximal subring of has the form or , we can apply Lemma 2.3, which yields . Since is -elementary, this means that either or . Because only has as a residue ring, we conclude that . ∎
Proposition 7.11.
Let be a -elementary ring that is not -elementary, and such that . Then, .
Proof.
By Lemma 7.6, has no residue ring isomorphic to . By Proposition 7.10, , and because is coverable by unital subrings. By Proposition 7.9, for all . Moreover, by Lemma 7.5(1), there is a unique pair with such that , and .
Thus, , which has unital covering number . By 3.2, occurs as a residue ring of . Thus, , which forces and . Since is -elementary, we must have . ∎
8. Bounds on the number of integers that are covering numbers of rings
This section is dedicated to the proof of Corollary 1.9. Recall that
Let denote the binary (base 2) logarithm. Our goal is to show that is bounded above by for some positive constant . Rather than attempting to find an optimal value for , we will be content with a value that is easy to verify.
Let be the prime counting function, which counts the number of primes less than or equal to a given positive number. The following bounds on will be useful in later estimates.
Lemma 8.1.
For all , . For all , .
Proof.
Let denote the natural logarithm. By [29, Corollary 1], we have when , and when . The remaining cases for the lower bound follow from inspection. ∎
Let be the prime power counting function, i.e.,
Using Lemma 8.1, it is not difficult to get an upper bound on .
Lemma 8.2.
Let . Then, .
Proof.
Let . Then,
∎
Using Lemma 8.2, we can provide bounds on the number of integers that are expressible in terms of the formulas given in the statement of Theorem 1.3.
Lemma 8.3.
Let be a natural number.
-
(1)
The number of integers , , that can be expressed as for some prime power is less than .
-
(2)
The number of integers , , that can be expressed as for some prime power is less than .
-
(3)
The number of integers , , that can be expressed as
where is some prime power, , is the smallest prime divisor of , and is the -binomial coefficient is less than .
-
(4)
Let be a natural number. The number of integers , , that can be expressed as
where and are prime powers, , , and is less than .
Proof.
The first two parts follow from the fact that there is one such integer of the desired form for each prime power. Hence, the number of such integers is bounded above by , which is less than by Lemma 8.2.
For (3), let
We first note that , from which we conclude that , and there at most choices for .
Next, by [33, Lemma 3.31] we have , so
It follows that and . Since there is exactly one integer expressible as for each pair , the number of such integers at most having this form is bounded above by
Finally, for (4), since , there are at most choices for , and, since and , there are at most choices for for a fixed , and at most choices for given a fixed . Hence, the number of integers expressible in the desired form is bounded above by the number of triples , which is bounded above by
∎
We now have what we need to prove Corollary 1.9.
References
- [1] Mira Bhargava. Groups as unions of proper subgroups. Amer. Math. Monthly, 116(5):413–422, 2009.
- [2] J. R. Britnell, A. Evseev, R. M. Guralnick, P. E. Holmes, and A. Maróti. Sets of elements that pairwise generate a linear group. J. Combin. Theory Ser. A, 115(3):442–465, 2008.
- [3] J. R. Britnell, A. Evseev, R. M. Guralnick, P. E. Holmes, and A. Maróti. Corrigendum to “Sets of elements that pairwise generate a linear group” [J. Combin. Theory Ser. A 115 (3) (2008) 442–465]. J. Combin. Theory Ser. A, 118(3):1152–1153, 2011.
- [4] R. A. Bryce, V. Fedri, and L. Serena. Subgroup coverings of some linear groups. Bull. Austral. Math. Soc., 60(2):227–238, 1999.
- [5] Merrick Cai and Nicholas J. Werner. Covering numbers of upper triangular matrix rings over finite fields. Involve, 12(6):1005–1013, 2019.
- [6] Pete L. Clark. Covering numbers in linear algebra. Amer. Math. Monthly, 119(1):65–67, 2012.
- [7] Jonathan Cohen. On rings as unions of four subrings. https://arxiv.org/abs/2008.03803, 2020.
- [8] J. H. E. Cohn. On -sum groups. Math. Scand., 75(1):44–58, 1994.
- [9] Eleonora Crestani. Sets of elements that pairwise generate a matrix ring. Comm. Algebra, 40(4):1570–1575, 2012.
- [10] Eloisa Detomi and Andrea Lucchini. On the structure of primitive -sum groups. Cubo, 10(3):195–210, 2008.
- [11] Casey Donoven and Luise-Charlotte Kappe. Finite coverings of semigroups and related structures. https://arxiv.org/pdf/2002.04072.pdf, 2020.
- [12] Stephen M. Gagola, III and Luise-Charlotte Kappe. On the covering number of loops. Expo. Math., 34(4):436–447, 2016.
- [13] Joseph A. Gallian. Contemporary abstract algebra. Cengage Learning, Ninth edition, 2017.
- [14] Martino Garonzi. Finite groups that are the union of at most 25 proper subgroups. J. Algebra Appl., 12(4):1350002, 11, 2013.
- [15] Martino Garonzi, Luise-Charlotte Kappe, and Eric Swartz. On integers that are covering numbers of groups. Experimental Mathematics, https://doi.org/10.1080/10586458.2019.1636425, 2019.
- [16] Soham Ghosh. The Zariski covering number for vector spaces and modules. Comm. Algebra, 50(5):1994–2017, 2022.
- [17] P. E. Holmes. Subgroup coverings of some sporadic groups. J. Combin. Theory Ser. A, 113(6):1204–1213, 2006.
- [18] Luise-Charlotte Kappe. Finite coverings: a journey through groups, loops, rings and semigroups. In Group theory, combinatorics, and computing, volume 611 of Contemp. Math., pages 79–88. Amer. Math. Soc., Providence, RI, 2014.
- [19] Apoorva Khare. Vector spaces as unions of proper subspaces. Linear Algebra Appl., 431(9):1681–1686, 2009.
- [20] Apoorva Khare and Akaki Tikaradze. Covering modules by proper submodules. Comm. Algebra, 50(2):498–507, 2022.
- [21] Serge Lang. Algebra, volume 211 of Graduate Texts in Mathematics. Springer-Verlag, New York, third edition, 2002.
- [22] Jacques Lewin. Subrings of finite index in finitely generated rings. J. Algebra, 5:84–88, 1967.
- [23] Andrea Lucchini and Attila Maróti. Rings as the unions of proper subrings. Algebr. Represent. Theory, 15(6):1035–1047, 2012.
- [24] Attila Maróti. Covering the symmetric groups with proper subgroups. J. Combin. Theory Ser. A, 110(1):97–111, 2005.
- [25] Bernard R. McDonald. Finite rings with identity. Pure and Applied Mathematics, Vol. 28. Marcel Dekker, Inc., New York, 1974.
- [26] B. H. Neumann. Groups covered by permutable subsets. J. London Math. Soc., 29:236–248, 1954.
- [27] G. Peruginelli and N. J. Werner. Maximal subrings and covering numbers of finite semisimple rings. Comm. Algebra, 46(11):4724–4738, 2018.
- [28] Richard S. Pierce. Associative algebras, volume 9 of Studies in the History of Modern Science. Springer-Verlag, New York-Berlin, 1982.
- [29] J. Barkley Rosser and Lowell Schoenfeld. Approximate formulas for some functions of prime numbers. Illinois J. Math., 6:64–94, 1962.
- [30] Gaetano Scorza. I gruppi che possone pensarsi come somma di tre lori sottogruppi. Boll. Un. Mat. Ital., 5:216–218, 1926.
- [31] Eric Swartz. On the covering number of symmetric groups having degree divisible by six. Discrete Math., 339(11):2593–2604, 2016.
- [32] Eric Swartz and Nicholas J. Werner. Covering numbers of commutative rings. J. Pure Appl. Algebra, 225(8):106622, 17, 2021.
- [33] Eric Swartz and Nicholas J. Werner. A new infinite family of -elementary rings. https://arxiv.org/abs/2211.10313, 2022.
- [34] M. J. Tomkinson. Groups as the union of proper subgroups. Math. Scand., 81(2):191–198, 1997.
- [35] Nicholas J. Werner. Covering numbers of finite rings. Amer. Math. Monthly, 122(6):552–566, 2015.