This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The covering numbers of rings

Eric Swartz Department of Mathematics, William & Mary, P.O. Box 8795, Williamsburg, VA 23187-8795, USA [email protected]  and  Nicholas J. Werner Department of Mathematics, Computer and Information Science, State University of New York College at Old Westbury, Old Westbury, NY 11560, USA [email protected]
Abstract.

A cover of an associative (not necessarily commutative nor unital) ring RR is a collection of proper subrings of RR whose set-theoretic union equals RR. If such a cover exists, then the covering number σ(R)\sigma(R) of RR is the cardinality of a minimal cover, and a ring RR is called σ\sigma-elementary if σ(R)<σ(R/I)\sigma(R)<\sigma(R/I) for every nonzero two-sided ideal II of RR. If RR is a ring with unity, then we define the unital covering number σu(R)\sigma_{u}(R) to be the size of a minimal cover of RR by subrings that contain 1R1_{R} (if such a cover exists), and RR is σu\sigma_{u}-elementary if σu(R)<σu(R/I)\sigma_{u}(R)<\sigma_{u}(R/I) for every nonzero two-sided ideal of RR. In this paper, we classify all σ\sigma-elementary unital rings and determine their covering numbers. Building on this classification, we are further able to classify all σu\sigma_{u}-elementary rings and prove σu(R)=σ(R)\sigma_{u}(R)=\sigma(R) for every σu\sigma_{u}-elementary ring RR. We also prove that, if RR is a ring without unity with a finite cover, then there exists a unital ring RR^{\prime} such that σ(R)=σu(R)\sigma(R)=\sigma_{u}(R^{\prime}), which in turn provides a complete list of all integers that are the covering number of a ring. Moreover, if

(N):={m:mN,σ(R)=m for some ring R},\mathscr{E}(N):=\{m:m\leqslant N,\sigma(R)=m\text{ for some ring }R\},

then we show that |(N)|=Θ(N/log(N))|\mathscr{E}(N)|=\Theta(N/\log(N)), 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 RR can be expressed as a union of proper subrings. Here, we use subring in the weakest sense: SRS\subseteq R is a subring when SS is an additive group and is closed under multiplication. In particular, SS need not contain a multiplicative identity, even if RR itself is unital. A cover of RR is a collection 𝒞\mathcal{C} of proper subrings of RR such that R=S𝒞SR=\bigcup_{S\in\mathcal{C}}S. When such a cover exists, we say that RR is coverable, and define the covering number σ(R)\sigma(R) to be the cardinality of a minimal cover. If RR is not coverable, we set σ(R)=\sigma(R)=\infty.

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 44-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 pn+1p^{n}+1 for every prime pp and positive integer nn 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 N129N\leqslant 129 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 n>2n>2, there exists a loop whose covering number is exactly nn, 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 n2n\geqslant 2, there exists an inverse semigroup whose covering number (by inverse subsemigroups) is exactly nn. 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 nn, 3n123\leqslant n\leqslant 12, examples are known of finite rings (with unity) that have covering number nn, 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 n>2n>2 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 σ(R)\sigma(R) for any ring RR 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 RR has a two-sided ideal II such that R/IR/I admits a cover 𝒞\mathcal{C}, then we may form a cover of RR by taking the inverse images of the subrings in 𝒞\mathcal{C} under the natural homomorphism RR/IR\to R/I. Thus, σ(R)σ(R/I)\sigma(R)\leqslant\sigma(R/I). Of capital importance are those rings for which the inequality σ(R)σ(R/I)\sigma(R)\leqslant\sigma(R/I) is strict for all I{0}I\neq\{0\}.

Definition 1.1.

A ring RR is said to be σ\sigma-elementary if σ(R)<σ(R/I)\sigma(R)<\sigma(R/I) for every nonzero two-sided ideal II of RR. Note that a σ\sigma-elementary ring RR must be coverable, since σ(R)<σ({0})=\sigma(R)<\sigma(\{0\})=\infty.

Observe that if σ(R)\sigma(R) is finite, then either RR is σ\sigma-elementary, or RR has a proper, σ\sigma-elementary residue ring with the same covering number. Moreover, if an integer nn occurs as the covering number of a ring, then there exists a σ\sigma-elementary ring RR such that σ(R)=n\sigma(R)=n. Thus, many questions about covering numbers of rings can be reduced to the case of σ\sigma-elementary rings.

In a recent paper [33], the authors introduced a new family of σ\sigma-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 σ\sigma-elementary rings with unity. Furthermore, we give formulas for the covering numbers of each family of σ\sigma-elementary rings.

In order to state Theorem 1.3, we must introduce some notation. For a prime power qq, 𝔽q\mathbb{F}_{q} is the finite field of order qq, and Mn(q)M_{n}(q) is the n×nn\times n matrix ring with entries from 𝔽q\mathbb{F}_{q}. The Jacobson radical of a ring RR is denoted by 𝒥(R)\mathscr{J}(R), and when RR contains unity, the unit group of RR is R×R^{\times}. When RR is a commutative ring and MM is an RR-module, the idealization of RR and MM, denoted R(+)MR(+)M, is the ring

R(+)M:={(r,m)rR,mM},R(+)M:=\{(r,m)\mid r\in R,m\in M\},

with multiplication (r1,m1)(r2,m2)=(r1r2,r1m2+r2m1)(r_{1},m_{1})\cdot(r_{2},m_{2})=(r_{1}r_{2},r_{1}m_{2}+r_{2}m_{1}). Of particular importance is the idealization 𝔽q(+)𝔽q2\mathbb{F}_{q}(+)\mathbb{F}_{q}^{2} of 𝔽q\mathbb{F}_{q} with the 2-dimensional vector space 𝔽q2\mathbb{F}_{q}^{2}. This ring can be represented as the following ring of 3×33\times 3 matrices:

𝔽q(+)𝔽q2{(abc0a000a):a,b,c𝔽q}.\mathbb{F}_{q}(+)\mathbb{F}_{q}^{2}\cong\left\{\begin{pmatrix}a&b&c\\ 0&a&0\\ 0&0&a\end{pmatrix}:a,b,c\in\mathbb{F}_{q}\right\}.

Next, we give some definitions from [33] related to rings of AGL-type.

Definition 1.2.

Given two powers q1q_{1} and q2q_{2} of a prime pp, we define q1q2q_{1}\otimes q_{2} to be the order of the field compositum of 𝔽q1\mathbb{F}_{q_{1}} and 𝔽q2\mathbb{F}_{q_{2}}, which is 𝔽q1𝔽p𝔽q2\mathbb{F}_{q_{1}}\otimes_{\mathbb{F}_{p}}\mathbb{F}_{q_{2}}. Observe that when q1=pd1q_{1}=p^{d_{1}} and q2=pd2q_{2}=p^{d_{2}}, we have q1q2=plcm(d1,d2)q_{1}\otimes q_{2}=p^{\text{lcm}(d_{1},d_{2})}.

Let n1n\geqslant 1 and let q=q1q2q=q_{1}\otimes q_{2}. We define A(n,q1,q2)A(n,q_{1},q_{2}) to be the following subring of Mn+1(q)M_{n+1}(q):

A(n,q1,q2):=(
Mn(q1) M×n1(q)
0 Fq2
 
)
,
A(n,q_{1},q_{2}):=\left(\text{\begin{tabular}[]{c|c}$M_{n}(q_{1})$&$M_{n\times 1}(q)$\\ \hline\cr$0$&$\mathbb{F}_{q_{2}}$\end{tabular} }\right),

and call this a ring of AGL-type. The construction and terminology for A(n,q1,q2)A(n,q_{1},q_{2}) were inspired by representations of the affine general linear group AGL(n,q){\rm AGL}(n,q), which is isomorphic to a subgroup of A(n,q,q)×A(n,q,q)^{\times}.

The formulas to compute σ(R)\sigma(R) require some accessory functions. For a prime power q=pdq=p^{d}, we take Irr(p,d){\rm Irr}(p,d) to be the set of all monic irreducible polynomials in 𝔽p[x]\mathbb{F}_{p}[x] of degree dd. We define

τ(q):={p, if d=1,|Irr(p,d)|+1, if d>1.\tau(q):=\begin{cases}p,&\text{ if }d=1,\\ |{\rm Irr}(p,d)|+1,&\text{ if }d>1.\end{cases}

We let ω\omega denote the prime omega function, which counts the number of distinct prime divisors of a natural number. Note that ω(1)=0\omega(1)=0. For q=pdq=p^{d}, we define

ν(q):={1, if d=1,ω(d), if d>1,\nu(q):=\begin{cases}1,&\text{ if }d=1,\\ \omega(d),&\text{ if }d>1,\end{cases}

which counts the number of maximal subrings of the field 𝔽q\mathbb{F}_{q}. The variation in the definitions between ω\omega and ν\nu is due to the fact that when q=pq=p, {0}\{0\} is the only maximal subring of 𝔽q\mathbb{F}_{q}; but, when q>pq>p, the maximal subrings of 𝔽q\mathbb{F}_{q} are exactly the maximal subfields of prime index. Finally, we let (nk)q\binom{n}{k}_{q} denote the qq-binomial coefficient, which counts the number of kk-dimensional subspaces of 𝔽qn\mathbb{F}_{q}^{n}:

(nk)q=(qn1)(qn11)(qn(k1)1)(qk1)(qk11)(q1).\binom{n}{k}_{q}=\frac{(q^{n}-1)(q^{n-1}-1)\cdots(q^{n-(k-1)}-1)}{(q^{k}-1)(q^{k-1}-1)\cdots(q-1)}.

We can now state our first main result.

Theorem 1.3.

Let RR be a σ\sigma-elementary ring with unity. Then, one of the following holds.

  1. (1)

    If RR is commutative and semisimple, then for some prime power q=pdq=p^{d}, Ri=1τ(q)𝔽qR\cong\bigoplus_{i=1}^{\tau(q)}\mathbb{F}_{q} and

    σ(R)=τ(q)ν(q)+d(τ(q)2).\sigma(R)=\tau(q)\nu(q)+d\binom{\tau(q)}{2}.
  2. (2)

    If RR is commutative but not semisimple, then for some prime power qq, R𝔽q(+)𝔽q2R\cong\mathbb{F}_{q}(+)\mathbb{F}_{q}^{2} and

    σ(R)=q+1.\sigma(R)=q+1.
  3. (3)

    If RR is noncommutative and semisimple, then for some prime power qq and integer n2n\geqslant 2, RMn(q)R\cong M_{n}(q) and

    σ(R)=1ak=1,akn1(qnqk)+k=1,akn/2(nk)q,\sigma(R)=\frac{1}{a}\prod_{k=1,\\ a\nmid k}^{n-1}(q^{n}-q^{k})+\sum_{k=1,\\ a\nmid k}^{\lfloor n/2\rfloor}{n\choose k}_{q},

    where aa is the smallest prime divisor of nn.

  4. (4)

    If RR is noncommutative and not semisimple, then RA(n,q1,q2)R\cong A(n,q_{1},q_{2}) and one of the two cases below holds. Let q=q1q2=q1dq=q_{1}\otimes q_{2}=q_{1}^{d}, and if n2n\geqslant 2, then let aa be the smallest prime divisor of nn.

    1. (i)

      n=1n=1 and (q1,q2)(2,2)(q_{1},q_{2})\neq(2,2) or (4,4)(4,4). In this case, σ(R)=q+1\sigma(R)=q+1.

    2. (ii)

      n3n\geqslant 3, d<n(n/a)d<n-(n/a), and (n,q1)(3,2)(n,q_{1})\neq(3,2). In this case,

      σ(R)=qn+(nd)q1+ω(d).\sigma(R)=q^{n}+\binom{n}{d}_{q_{1}}+\omega(d).

Portions of Theorem 1.3 have been proved in earlier papers. The classification of commutative unital σ\sigma-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 Mn(q)M_{n}(q) is due to Crestani, Lucchini, and Maróti [9, 23]. Since Mn(q)M_{n}(q) is simple, it is clearly σ\sigma-elementary, and the fact that these are the only noncommutative semisimple σ\sigma-elementary rings follows from the classification of covering numbers for finite semisimple rings done in [27]. The determination of when A(n,q1,q2)A(n,q_{1},q_{2}) is σ\sigma-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 σ\sigma-elementary unital ring falls into one of the four classes listed in Theorem 1.3.

While Theorem 1.3 handles the situation where RR 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 RR in which every subring contains 1R1_{R}.

Definition 1.4.

Let RR be a ring with unity. If RR can be covered by unital subrings (i.e. those containing 1R1_{R}), then the unital covering number σu(R)\sigma_{u}(R) is defined to be the cardinality of a minimal cover by unital subrings. We say RR is σu\sigma_{u}-elementary if σu(R)<σu(R/I)\sigma_{u}(R)<\sigma_{u}(R/I) for every nonzero two-sided ideal II of RR. Note that a σu\sigma_{u}-elementary ring necessarily contains unity, and admits a cover by unital subrings.

Theorem 1.5.

Let RR be a ring without unity that has a finite cover. Then, there exists a ring RR^{\prime} with unity such that σ(R)=σu(R)\sigma(R)=\sigma_{u}(R^{\prime}). 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 σu\sigma_{u}-elementary rings. Our classification is very similar to that of unital σ\sigma-elementary rings, which was given in Theorem 1.3.

Theorem 1.6.

Let RR be a σu\sigma_{u}-elementary ring.

  • (1)

    If RR is σ\sigma-elementary, then RR is one of the rings listed in Theorem 1.3, with the exception that R≇i=1p𝔽pR\not\cong\bigoplus_{i=1}^{p}\mathbb{F}_{p}. Moreover, any ring R≇i=1p𝔽pR\not\cong\bigoplus_{i=1}^{p}\mathbb{F}_{p} listed in Theorem 1.3 is σu\sigma_{u}-elementary.

  • (2)

    If RR is not σ\sigma-elementary, then either Ri=1p+1𝔽pR\cong\bigoplus_{i=1}^{p+1}\mathbb{F}_{p} with σu(R)=p+(p2)\sigma_{u}(R)=p+\binom{p}{2}, or RA(1,2,2)R\cong A(1,2,2) with σu(R)=3\sigma_{u}(R)=3.

In particular, if RR is σu\sigma_{u}-elementary, then σu(R)=σ(R)\sigma_{u}(R)=\sigma(R), 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 1313.

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 (N):={m:mN,σ(R)=m for some ring R}.\mathscr{E}(N):=\{m:m\leqslant N,\sigma(R)=m\text{ for some ring }R\}. Then, for all N5N\geqslant 5,

NlogN<|(N)|<144NlogN,\frac{N}{\log N}<|\mathscr{E}(N)|<\frac{144N}{\log N},

where logN\log N denotes the binary (base 22) logarithm of NN. In particular, we have |(N)|=Θ(N/log(N))|\mathscr{E}(N)|=\Theta(N/\log(N)) and

limN|(N)|N=0.\lim_{N\to\infty}\frac{|\mathscr{E}(N)|}{N}=0.

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 σ\sigma-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 σ\sigma-elementary ring must in fact be {0}\{0\} (so, a σ\sigma-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 σ\sigma-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 σ(R)\sigma(R) 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 RR be a ring with unity.

  1. (1)

    RR is coverable if and only if RR cannot be generated (as a ring) by a single element.

  2. (2)

    If RR is noncommutative, then RR is coverable.

  3. (3)

    For any two-sided ideal II of RR, a cover of R/IR/I can be lifted to a cover of RR. Hence, σ(R)σ(R/I)\sigma(R)\leqslant\sigma(R/I).

  4. (4)

    If each proper subring of RR is contained in a maximal subring, then we may assume that any minimal cover of RR consists of maximal subrings.

By Lemma 2.1(4), maximal subrings of RR 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.

Let RR be a coverable ring, and let 𝒞\mathcal{C} be a minimal cover of RR.

  1. (1)

    [35, Lemma 2.1] If MM is a maximal subring of RR and M𝒞M\notin\mathcal{C}, then σ(M)σ(R)\sigma(M)\leq\sigma(R).

  2. (2)

    [32, Lemma 2.2] Let SS be a proper subring of RR such that R=SIR=S\oplus I for some two-sided ideal II of RR. If σ(R)<σ(R/I)\sigma(R)<\sigma(R/I), then STS\subseteq T for some T𝒞T\in\mathcal{C}. If, in addition, SS is a maximal subring of RR, then S𝒞S\in\mathcal{C}.

As shown in [35, Theorem 2.2], when RR is a direct sum of rings, σ(R)\sigma(R) can be determined from the covering numbers of the summands of RR as long as all maximal subrings of RR respect the direct sum decomposition. While not explicitly stated in [35], the analogous result for unital maximal subrings and unital coverings of RR is also true, and can be proved in the same manner as [35, Theorem 2.2].

Lemma 2.3.

[35, Theorem 2.2] Let R=i=1tRiR=\bigoplus_{i=1}^{t}R_{i} for rings R1,,RtR_{1},\ldots,R_{t}. Assume that each maximal subring MM of RR has the form

M=R1Ri1MiRi+1RtM=R_{1}\oplus\cdots\oplus R_{i-1}\oplus M_{i}\oplus R_{i+1}\oplus\cdots\oplus R_{t}

for some 1it1\leqslant i\leqslant t. Then, σ(R)=min1it{σ(Ri)}\sigma(R)=\min_{1\leqslant i\leqslant t}\{\sigma(R_{i})\}. Furthermore, if each RiR_{i} is a unital ring and each unital maximal subring has the above form (with MiM_{i} itself unital), then σu(R)=min1it{σu(Ri)}\sigma_{u}(R)=\min_{1\leqslant i\leqslant t}\{\sigma_{u}(R_{i})\}.

From results of Neumann [26, Lemma 4.1, 4.4] and Lewin [22, Lemma 1], it is known that if RR has a finite covering number, then there exists a two-sided ideal II of RR of finite index such that σ(R)=σ(R/I)\sigma(R)=\sigma(R/I). Hence, the calculation of covering numbers reduces to the case of finite rings. In fact, a much stronger reduction is possible. Recall that 𝒥(R)\mathscr{J}(R) denotes the Jacobson radical of a ring RR.

Lemma 2.4.

[32, Theorem 3.12] Let RR be a (unital, associative) ring such that σ(R)\sigma(R) is finite. Then, there exists a two-sided ideal II of RR such that R/IR/I is finite; R/IR/I has characteristic pp; 𝒥(R/I)2={0}\mathscr{J}(R/I)^{2}=\{0\}; and σ(R/I)=σ(R)\sigma(R/I)=\sigma(R).

Thus, for unital rings with finite covers, the computation of σ(R)\sigma(R) reduces to the case where RR is finite of characteristic pp and 𝒥(R)2={0}\mathscr{J}(R)^{2}=\{0\}. When RR is finite, R/𝒥(R)R/\mathscr{J}(R) 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 RR, and how it relates to R/𝒥(R)R/\mathscr{J}(R).

Theorem 2.5.

(Wedderburn-Malcev Theorem) Let RR be a finite ring with unity of characteristic pp. Then, there exists an 𝔽p\mathbb{F}_{p}-subalgebra SS of RR such that R=S𝒥(R)R=S\oplus\mathscr{J}(R), and SR/𝒥(R)S\cong R/\mathscr{J}(R) as 𝔽p\mathbb{F}_{p}-algebras. Moreover, SS is unique up to conjugation by elements of 1+𝒥(R)1+\mathscr{J}(R).

Let 𝒮(R)\mathscr{S}(R) be the set of all semisimple complements to 𝒥(R)\mathscr{J}(R) in RR. By Theorem 2.5, all such complements are conjugate. So, for any S𝒮(R)S\in\mathscr{S}(R), we have R=S𝒥(R)R=S\oplus\mathscr{J}(R), and SR/𝒥(R)S\cong R/\mathscr{J}(R).

The decomposition given by Theorem 2.5 allows us to characterize all of the maximal subrings of RR.

Lemma 2.6.

[32, Theorem 3.10] Let RR be a finite ring with unity of characteristic pp, let MM be a maximal subring of RR, and let J=𝒥(R)J=\mathscr{J}(R).

  1. (1)

    JMJ\subseteq M if and only if MM is the inverse image of a maximal subring of R/JR/J.

  2. (2)

    JMJ\not\subseteq M if and only if M=S𝒥(M)M=S\oplus\mathscr{J}(M), where S𝒮(R)S\in\mathscr{S}(R) and 𝒥(M)=MJ\mathscr{J}(M)=M\cap J is an ideal of RR that is maximal among the subideals of RR contained in JJ.

3. Peirce decomposition and bounds

We now turn our attention to proving that any σ\sigma-elementary unital ring falls into one of the four classes given in Theorem 1.3. Throughout this section, we will assume that RR is a σ\sigma-elementary unital ring with σ(R)\sigma(R) finite. By Lemma 2.4, we can assume that RR is finite of characteristic pp and has Jacobson radical J=𝒥(R)J=\mathscr{J}(R) with J2={0}J^{2}=\{0\}. Moreover, using Theorem 2.5, we have R=SJR=S\oplus J, where SR/JS\cong R/J is a semisimple complement to JJ in RR. The set of all such complements is denoted by 𝒮(R)\mathscr{S}(R). For the remainder of this section, we fix S𝒮(R)S\in\mathscr{S}(R). Then, for some N1N\geqslant 1, S=i=1NSiS=\bigoplus_{i=1}^{N}S_{i}, where each SiS_{i} is a simple ring. For each 1iN1\leqslant i\leqslant N, let Si=Mni(qi)S_{i}=M_{n_{i}}(q_{i}), where ni1n_{i}\geqslant 1 and qiq_{i} is a power of pp. Finally, let ei=1Sie_{i}=1_{S_{i}}, so that e1,,eNe_{1},\ldots,e_{N} are orthogonal idempotents such that i=1Nei\sum_{i=1}^{N}e_{i} is the unity of RR.

Apply a two-sided Peirce decomposition to JJ using the idempotents e1,,eNe_{1},\ldots,e_{N}. This gives J=1i,jNeiJejJ=\bigoplus_{1\leqslant i,j\leqslant N}e_{i}Je_{j}, where each eiJeje_{i}Je_{j} is an (SiS_{i}, SjS_{j})-bimodule. If Si=Mni(Fi)S_{i}=M_{n_{i}}(F_{i}) and Sj=Mnj(Fj)S_{j}=M_{n_{j}}(F_{j}), then (SiS_{i}, SjS_{j})-bimodules are equivalent to modules over Mninj(F)M_{n_{i}n_{j}}(F), where F=Fi𝔽pFjF=F_{i}\otimes_{\mathbb{F}_{p}}F_{j} is the compositum of FiF_{i} and FjF_{j}. In particular, we may speak of the length of eiJeje_{i}Je_{j} as an (SiS_{i}, SjS_{j})-bimodule, i.e., the number of simple (SiS_{i}, SjS_{j})-bimodules in a direct sum decomposition of eiJeje_{i}Je_{j}. Note that any simple (SiS_{i}, SjS_{j})-bimodule has order |F|ninj|F|^{n_{i}n_{j}}.

Definition 3.1.

For all 1i,jN1\leqslant i,j\leqslant N, let Jij:=eiJejJ_{ij}:=e_{i}Je_{j}. If Jij{0}J_{ij}\neq\{0\}, then let λij\lambda_{ij} be the length of JijJ_{ij} as an (SiS_{i}, SjS_{j})-bimodule. If Jij={0}J_{ij}=\{0\}, then we take λij=0\lambda_{ij}=0.

Recall that when qiq_{i} and qjq_{j} are powers of pp, we define qiqjq_{i}\otimes q_{j} to be the order of 𝔽qi𝔽p𝔽qj\mathbb{F}_{q_{i}}\otimes_{\mathbb{F}_{p}}\mathbb{F}_{q_{j}}. If qi=pdiq_{i}=p^{d_{i}} and qj=pdjq_{j}=p^{d_{j}}, then qiqj=plcm(di,dj)q_{i}\otimes q_{j}=p^{\text{lcm}(d_{i},d_{j})}. Note also that we have Mni(qi)𝔽pMnj(qj)Mninj(qiqj)M_{n_{i}}(q_{i})\otimes_{\mathbb{F}_{p}}M_{n_{j}}(q_{j})\cong M_{n_{i}n_{j}}(q_{i}\otimes q_{j}) for any positive integers nin_{i} and njn_{j}. Thus, |Jij|=(qiqj)ninjλij|J_{ij}|=(q_{i}\otimes q_{j})^{n_{i}n_{j}\lambda_{ij}}.

We remark that the rings in parts (2) and (4) of Theorem 1.3 may occur as subrings of RR, and can be described using the notation and parameters we have defined. If nk=1n_{k}=1 and λkk=2\lambda_{kk}=2 for some 1kN1\leqslant k\leqslant N, then SkJkk𝔽qk(+)𝔽qk2S_{k}\oplus J_{kk}\cong\mathbb{F}_{q_{k}}(+)\mathbb{F}_{q_{k}}^{2}. If 1i,jN1\leqslant i,j\leqslant N with iji\neq j, nj=1n_{j}=1, and λij=1\lambda_{ij}=1, then SiSjJijA(ni,qi,qj)S_{i}\oplus S_{j}\oplus J_{ij}\cong A(n_{i},q_{i},q_{j}) is a ring of AGL-type.

As we now show, any (SiS_{i}, SjS_{j})-bimodule inside of JJ is a two-sided ideal of RR, and has an ideal complement in JJ.

Lemma 3.2.

Let 1i,jN1\leqslant i,j\leqslant N.

  1. (1)

    Let XJX\subseteq J be an (SiS_{i}, SjS_{j})-bimodule. Then, XJijX\subseteq J_{ij}, XX is a two-sided ideal of RR, and XX has a two-sided ideal complement in JJ.

  2. (2)

    Let IJI\subseteq J be a two-sided ideal of RR. Then, II has a two-sided ideal complement in JJ.

  3. (3)

    Let IJI\subseteq J be a two-sided ideal of RR and let Z=JZ(R)Z=J\cap Z(R). For all xJx\in J, SI=S1+xIS\oplus I=S^{1+x}\oplus I if and only if xI+Zx\in I+Z. Hence, the number of conjugates of SIS\oplus I in RR is equal to |J:(I+Z)||J:(I+Z)|.

  4. (4)

    For each 1iN1\leqslant i\leqslant N, RR has a residue ring isomorphic to SiJiiS_{i}\oplus J_{ii}. For all 1i,jN1\leqslant i,j\leqslant N with iji\neq j, RR has a residue ring isomorphic to SiSjJijS_{i}\oplus S_{j}\oplus J_{ij}.

Proof.

(1) Since eie_{i} and eje_{j} are the identities of SiS_{i} and SjS_{j}, respectively, X=eiXejJijX=e_{i}Xe_{j}\subseteq J_{ij}. To show that XX is a two-sided ideal of RR, let rRr\in R and xXx\in X. Then, r=s+yr=s+y for some sSs\in S and yJy\in J. Since J2={0}J^{2}=\{0\}, rx=sxrx=sx. Next, write s=k=1Nsks=\sum_{k=1}^{N}s_{k}, where each skSks_{k}\in S_{k}. Since the idempotents e1,,eNe_{1},\ldots,e_{N} are orthogonal, we have sei=sieise_{i}=s_{i}e_{i}. So, rx=sx=s(eix)=sixXrx=sx=s(e_{i}x)=s_{i}x\in X. The proof that xrXxr\in X is similar. Finally, since SiS_{i} and SjS_{j} are simple rings, JJ is a semisimple (SiS_{i}, SjS_{j})-bimodule. So, there is an (SiS_{i}, SjS_{j})-bimodule X^J\widehat{X}\subseteq J such that J=XX^J=X\oplus\widehat{X}, and by the work above, X^\widehat{X} is also a two-sided ideal of RR.

(2) Apply the Peirce decomposition to II to get I=i,jeiIejI=\bigoplus_{i,j}e_{i}Ie_{j}. Then, each eiIeje_{i}Ie_{j} is an (SiS_{i}, SjS_{j})-bimodule contained in JJ. By part (1), each eiIeje_{i}Ie_{j} has an ideal complement, and therefore II does as well.

(3) Note that since J2={0}J^{2}=\{0\}, (1+x)1=1x(1+x)^{-1}=1-x for all xJx\in J. So, s1+x=(1x)s(1+x)=s+sxxss^{1+x}=(1-x)s(1+x)=s+sx-xs for all sSs\in S and all xJx\in J. Certainly, if xI+Zx\in I+Z, then S1+xSIS^{1+x}\subseteq S\oplus I. Conversely, assume that xJx\in J is such that SI=S1+xIS\oplus I=S^{1+x}\oplus I. By (2), there is an ideal complement I^\widehat{I} of II in JJ. Express xx as x=α+βx=\alpha+\beta, where αI\alpha\in I and βI^\beta\in\widehat{I}. Then, for all sSs\in S,

sxxs=(sααs)+(sββs),sx-xs=(s\alpha-\alpha s)+(s\beta-\beta s),

and sxxsIsx-xs\in I because s1+xSIs^{1+x}\in S\oplus I. It follows that sββsII^={0}s\beta-\beta s\in I\cap\widehat{I}=\{0\}. Thus, β\beta is central in RR and xI+Zx\in I+Z.

(4) We will prove the claim regarding SiSjJijS_{i}\oplus S_{j}\oplus J_{ij}. The proof of the other statement is similar. Let T=SiSjJijT=S_{i}\oplus S_{j}\oplus J_{ij}, and let S^ij=ki,kjSk\widehat{S}_{ij}=\bigoplus_{k\neq i,k\neq j}S_{k}, which is a direct sum complement to SiSjS_{i}\oplus S_{j} in SS. Let J^ij\widehat{J}_{ij} be an ideal complement to JijJ_{ij} in JJ, which exists by part (2). Since the idempotents e1,,eNe_{1},\ldots,e_{N} are orthogonal, S^ij\widehat{S}_{ij} and TT annihilate one another. Hence, S^ijJ^ij\widehat{S}_{ij}\oplus\widehat{J}_{ij} is a two-sided ideal of RR, and R/(S^ijJ^ij)TR/(\widehat{S}_{ij}\oplus\widehat{J}_{ij})\cong T. ∎

Proposition 3.3.

Assume that RR is σ\sigma-elementary. Let MM be a maximal subring of RR that does not contain JJ. Then, MM is contained in every minimal cover of RR.

Proof.

By Lemma 2.6, M=SIM=S^{\prime}\oplus I, where S𝒮(R)S^{\prime}\in\mathscr{S}(R) and II is a maximal subideal of JJ. By Lemma 3.2, II has an ideal complement I^\widehat{I} in JJ. Thus, R=MI^R=M\oplus\widehat{I}. Since RR is σ\sigma-elementary, MM must be contained in every minimal cover of RR by Lemma 2.2. ∎

Proposition 3.3 (often in conjunction with Lemma 2.2) will be used to produce lower bounds on the covering number of a σ\sigma-elementary ring. The remainder of this section contains various bounds on the covering numbers of RR and its subrings.

Lemma 3.4.

Let 1i,jN1\leqslant i,j\leqslant N with iji\neq j, and let T=SiSjJijT=S_{i}\oplus S_{j}\oplus J_{ij}.

  1. (1)

    JijZ(T)={0}J_{ij}\cap Z(T)=\{0\}.

  2. (2)

    Let II be a subideal of JijJ_{ij}. Then, for all xJijx\in J_{ij}, (SiSj)I=(SiSj)1+xI(S_{i}\oplus S_{j})\oplus I=(S_{i}\oplus S_{j})^{1+x}\oplus I if and only if xIx\in I.

  3. (3)

    σ(R)σ(T)\sigma(R)\leqslant\sigma(T).

Proof.

For (1), there is nothing to prove if Jij={0}J_{ij}=\{0\}, so assume there exists xJij{0}x\in J_{ij}\setminus\{0\}. Then, eix=xe_{i}x=x and xei=(xej)ei=0xe_{i}=(xe_{j})e_{i}=0. Hence, xZ(T)x\in Z(T) if and only if x=0x=0. Part (2) now follows from part (1) and Lemma 3.2(3), and part (3) follows from Lemma 3.2(4). ∎

The next two lemmas are generalizations of [33, Proposition 3.3].

Lemma 3.5.

Let 1i,jN1\leqslant i,j\leqslant N and let q=qiqjq=q_{i}\otimes q_{j}. Assume that Jij{0}J_{ij}\neq\{0\}, and let T=SiSjJijT=S_{i}\oplus S_{j}\oplus J_{ij}. Then, σ(T)(qninj+11)/(q1)\sigma(T)\leqslant(q^{n_{i}n_{j}+1}-1)/(q-1).

Proof.

For convenience, let S=SiSjS=S_{i}\oplus S_{j} and J=JijJ=J_{ij}. First we argue that we may assume JJ is a simple (SiS_{i}, SjS_{j})-bimodule. Indeed, if JJ is not simple, then let XJX\subseteq J be a simple (SiS_{i}, SjS_{j})-bimodule. By Lemma 3.2, XX has an ideal complement X^\widehat{X} in JJ. Since T/X^SiSjXT/\widehat{X}\cong S_{i}\oplus S_{j}\oplus X, we get σ(T)σ(T/X^)=σ(SiSjX)\sigma(T)\leqslant\sigma(T/\widehat{X})=\sigma(S_{i}\oplus S_{j}\oplus X). Hence, we may assume that JJ is simple.

For each nonzero xJx\in J, define CS(x):={sS:sx=xs}C_{S}(x):=\{s\in S:sx=xs\}. Then, each CS(x)C_{S}(x) is a subring of SS. By Lemma 3.4(1), CS(x)SC_{S}(x)\subsetneq S for all nonzero xJx\in J, so CS(x)JC_{S}(x)\oplus J is a proper subring of TT for all x0x\neq 0.

Let 𝒞=𝒮(T){CS(x)J:xJ,x0}\mathcal{C}=\mathscr{S}(T)\cup\{C_{S}(x)\oplus J:x\in J,x\neq 0\}. We will show that 𝒞\mathcal{C} is a cover of TT. Suppose tT\xJ{0}CS(x)Jt\in T\backslash\bigcup_{x\in J\setminus\{0\}}C_{S}(x)\oplus J; that is, suppose ts+Jt\in s+J for some sS\xJ{0}CS(x)s\in S\backslash\bigcup_{x\in J\setminus\{0\}}C_{S}(x).

We claim that s+J=s1+Js+J=s^{1+J}. We know that s1+x=s+sxxss^{1+x}=s+sx-xs for all xJx\in J, so s1+Js+Js^{1+J}\subseteq s+J. Furthermore, s1+x1=s1+x2s^{1+x_{1}}=s^{1+x_{2}} if and only if s1+(x1x2)=ss^{1+(x_{1}-x_{2})}=s. Since sCS(x)s\notin C_{S}(x) for all x0x\neq 0, this means that s1+x1=s1+x2s^{1+x_{1}}=s^{1+x_{2}} if and only if x1=x2x_{1}=x_{2}. Hence, |s1+J|=|J|=|s+J||s^{1+J}|=|J|=|s+J|, and so s+J=s1+Js+J=s^{1+J}. Thus,

ts+J=s1+JS𝒮(T)S,t\in s+J=s^{1+J}\subseteq\bigcup_{S^{\prime}\in\mathscr{S}(T)}S^{\prime},

and 𝒞\mathcal{C} is a cover of TT.

Finally, note that CS(cx)=CS(x)C_{S}(cx)=C_{S}(x) whenever c𝔽q×c\in\mathbb{F}_{q}^{\times}. This means that there are at most (|J|1)/(q1)=(qninj1)/(q1)(|J|-1)/(q-1)=(q^{n_{i}n_{j}}-1)/(q-1) subrings CS(x)C_{S}(x) of SS. Since JJ is a simple (Si,Sj)(S_{i},S_{j})-bimodule, |J|=qninj|J|=q^{n_{i}n_{j}}. Therefore,

σ(T)|J|+|J|1q1=qninj+11q1,\sigma(T)\leqslant|J|+\frac{|J|-1}{q-1}=\frac{q^{n_{i}n_{j}+1}-1}{q-1},

as required. ∎

Lemma 3.6.

Let 1i,jN1\leqslant i,j\leqslant N such that iji\neq j and nj=1n_{j}=1, and let q=qiqjq=q_{i}\otimes q_{j}. If JijJ_{ij} is a simple (SiS_{i}, SjS_{j})-bimodule and RR is σ\sigma-elementary, then qni+1σ(R)q^{n_{i}}+1\leqslant\sigma(R). If, in addition, ni=1n_{i}=1, then R=SiSjJijR=S_{i}\oplus S_{j}\oplus J_{ij} and σ(R)=q+1\sigma(R)=q+1.

Proof.

Assume that JijJ_{ij} is simple and RR is σ\sigma-elementary. Let T=SiSjJijT=S_{i}\oplus S_{j}\oplus J_{ij}. Then, for each xJijx\in J_{ij}, (SiSj)1+x(S_{i}\oplus S_{j})^{1+x} is a maximal subring of TT. By Lemma 3.4, JijZ(T)={0}J_{ij}\cap Z(T)=\{0\}, so there are |Jij|=qni|J_{ij}|=q^{n_{i}} distinct conjugates of SiSjS_{i}\oplus S_{j} in TT.

Let S^ij=ki,kjSk\widehat{S}_{ij}=\bigoplus_{k\neq i,k\neq j}S_{k} and let J^ij\widehat{J}_{ij} be an ideal complement to JijJ_{ij} in JJ. Then, for any conjugate SS^{\prime} of SiSjS_{i}\oplus S_{j} in TT, we have

R=(SS^ijJ^ij)Jij.R=(S^{\prime}\oplus\widehat{S}_{ij}\oplus\widehat{J}_{ij})\oplus J_{ij}.

Since JijJ_{ij} is simple, each subring SS^ijJ^ijS^{\prime}\oplus\widehat{S}_{ij}\oplus\widehat{J}_{ij} is maximal. On the one hand, because RR is σ\sigma-elementary, each such subring is contained in every minimal cover of RR by Lemma 2.2. So, qniσ(R)q^{n_{i}}\leqslant\sigma(R). On the other hand, the union of all the conjugates SS^{\prime} cannot cover TT. To see this, let ATA\subseteq T be the intersection of all qniq^{n_{i}} conjugates of SiSjS_{i}\oplus S_{j}. Then, the union of all the conjugates covers at most

|Jij|(|SiSj||A|)+|A|=|T||A|(|J|1)|J_{ij}|\big{(}|S_{i}\oplus S_{j}|-|A|\big{)}+|A|=|T|-|A|\big{(}|J|-1\big{)}

elements of TT. It follows that the union of all the subrings SS^ijJ^ijS^{\prime}\oplus\widehat{S}_{ij}\oplus\widehat{J}_{ij} cannot cover RR. Hence, qni+1σ(R)q^{n_{i}}+1\leqslant\sigma(R).

Lastly, by Lemmas 3.4 and 3.5, σ(R)σ(T)(qni+11)/(q1)\sigma(R)\leqslant\sigma(T)\leqslant(q^{n_{i}+1}-1)/(q-1). When ni=1n_{i}=1, this forces σ(R)=q+1\sigma(R)=q+1 and R=TR=T because RR is σ\sigma-elementary. ∎

Recall the definitions for Irr(p,d)\text{Irr}(p,d) and τ(q)\tau(q) stated prior to Theorem 1.3. It is shown in [35, Theorem 3.5] that a direct sum i=1t𝔽q\bigoplus_{i=1}^{t}\mathbb{F}_{q} of copies of 𝔽q\mathbb{F}_{q} is coverable if and only if tτ(q)t\geqslant\tau(q).

Lemma 3.7.

Let q=pdq=p^{d}, let tτ(q)t\geqslant\tau(q), and let T=i=1t𝔽qT=\bigoplus_{i=1}^{t}\mathbb{F}_{q}. If d=1d=1, then σ(T)=12(p2+p)\sigma(T)=\tfrac{1}{2}(p^{2}+p). If d2d\geqslant 2, then σ(T)q2/(2d)\sigma(T)\leqslant q^{2}/(2d).

Proof.

As stated in Theorem 1.3(1), we have

σ(T)=τ(q)ν(q)+d(τ(q)2),\sigma(T)=\tau(q)\nu(q)+d\displaystyle\binom{\tau(q)}{2},

where ν(q)\nu(q) is equal to 1 if d=1d=1, and is equal to the number of prime divisors of dd otherwise. Via this formula, it is clear that σ(T)=p+(p2)=12(p2+p)\sigma(T)=p+\binom{p}{2}=\tfrac{1}{2}(p^{2}+p) when d=1d=1. So, assume that d2d\geqslant 2. Then, ν(q)d2\nu(q)\leqslant\tfrac{d}{2}, and

(3.8) σ(T)τ(q)d2+d2τ(q)(τ(q)1)=d2τ(q)2.\sigma(T)\leqslant\tau(q)\cdot\tfrac{d}{2}+\tfrac{d}{2}\cdot\tau(q)\cdot(\tau(q)-1)=\tfrac{d}{2}\cdot\tau(q)^{2}.

Now, it is well known that |Irr(p,d)|pd/d|\text{Irr}(p,d)|\leqslant p^{d}/d, with equality if and only if d=1d=1. When d2d\geqslant 2, |Irr(p,d)|<pd/d|\text{Irr}(p,d)|<p^{d}/d, and hence τ(q)q/d\tau(q)\leqslant q/d. Combining this with (3.8) produces the desired upper bound for σ(T)\sigma(T). ∎

4. Restrictions on JJ

Maintain the notation given at the start of Section 3. Our goal in this section is to prove that when RR is σ\sigma-elementary, most of the bimodules JijJ_{ij} must be zero. In essence, this means that a σ\sigma-elementary ring is “almost” semisimple. We will first examine the bimodules JijJ_{ij} with iji\neq j, and then consider the (SkS_{k}, SkS_{k})-bimodules JkkJ_{kk}.

Theorem 4.1.

Let 1i,jN1\leqslant i,j\leqslant N with iji\neq j. Assume that Jij{0}J_{ij}\neq\{0\} and RR is σ\sigma-elementary. Then, the following hold:

  1. (1)

    Jk={0}J_{k\ell}=\{0\} for all 1k,N1\leqslant k,\ell\leqslant N such that kk\neq\ell and (k,)(i,j)(k,\ell)\neq(i,j).

  2. (2)

    JijJ_{ij} is a simple (SiS_{i}, SjS_{j})-bimodule.

  3. (3)

    At least one of SiS_{i} or SjS_{j} is a field.

Proof.

Throughout, let Tij=SiSjJijT_{ij}=S_{i}\oplus S_{j}\oplus J_{ij} and q=qiqjq=q_{i}\otimes q_{j}.

(1) Let IijI_{ij} be a maximal subideal of JijJ_{ij}. If SS^{\prime} is a conjugate of SiSjS_{i}\oplus S_{j} in TijT_{ij}, then SIijS^{\prime}\oplus I_{ij} is a maximal subring of TijT_{ij}. Let J^ij\widehat{J}_{ij} be an ideal complement to JijJ_{ij} in JJ, and let

S^ij:=mi,mjSm,\widehat{S}_{ij}:=\bigoplus_{m\neq i,m\neq j}S_{m},

which is a direct sum complement to SiSjS_{i}\oplus S_{j} in SS. Then, for any conjugate SS^{\prime} of SiSjS_{i}\oplus S_{j} in TijT_{ij},

(4.2) (SS^ij)(IijJ^ij)(S^{\prime}\oplus\widehat{S}_{ij})\oplus(I_{ij}\oplus\widehat{J}_{ij})

is a maximal subring of RR. The number of such subrings is equal to the number of distinct maximal subrings of TijT_{ij} of the form SIijS^{\prime}\oplus I_{ij}. Since the size of a simple (Si,Sj)(S_{i},S_{j})-bimodule is qninjq^{n_{i}n_{j}}, by Lemma 3.4(2), RR contains |Jij:Iij|=qninj|J_{ij}:I_{ij}|=q^{n_{i}n_{j}} maximal subrings of the form in (4.2).

Suppose that Jk{0}J_{k\ell}\neq\{0\} for some (k,)(i,j)(k,\ell)\neq(i,j). The arguments of the previous paragraph can be applied to SkSJkS_{k}\oplus S_{\ell}\oplus J_{k\ell}. Employing the analogous notation and letting q=qkqq^{\prime}=q_{k}\otimes q_{\ell}, we see that RR contains (q)nkn(q^{\prime})^{n_{k}n_{\ell}} maximal subrings of the form

(4.3) (S′′S^k)(IkJ^k),(S^{\prime\prime}\oplus\widehat{S}_{k\ell})\oplus(I_{k\ell}\oplus\widehat{J}_{k\ell}),

where S′′S^{\prime\prime} is a conjugate of SkSS_{k}\oplus S_{\ell}. Moreover, no ring of the form of (4.2) can equal a ring of the form of (4.3), because IijJ^ijIkJ^kI_{ij}\oplus\widehat{J}_{ij}\neq I_{k\ell}\oplus\widehat{J}_{k\ell}. By Proposition 3.3,

(4.4) qninj+(q)nknσ(R).q^{n_{i}n_{j}}+(q^{\prime})^{n_{k}n_{\ell}}\leqslant\sigma(R).

Now, by Lemmas 3.4 and 3.5,

(4.5) σ(R)<σ(Tij)qninj+11q1.\sigma(R)<\sigma(T_{ij})\leqslant\dfrac{q^{n_{i}n_{j}+1}-1}{q-1}.

Thus, by (4.4) and (4.5), qninj+(q)nkn<(qninj+11)/(q1)q^{n_{i}n_{j}}+(q^{\prime})^{n_{k}n_{\ell}}<(q^{n_{i}n_{j}+1}-1)/(q-1). This gives

(q)nkn<qninj+11q1qninj=qninj1q1<qninj.(q^{\prime})^{n_{k}n_{\ell}}<\dfrac{q^{n_{i}n_{j}+1}-1}{q-1}-q^{n_{i}n_{j}}=\dfrac{q^{n_{i}n_{j}}-1}{q-1}<q^{n_{i}n_{j}}.

However, we also have σ(R)<σ(SkSJk)\sigma(R)<\sigma(S_{k}\oplus S_{\ell}\oplus J_{k\ell}). Working as above, this leads to qninj<(q)nknq^{n_{i}n_{j}}<(q^{\prime})^{n_{k}n_{\ell}}, a contradiction.

(2) Suppose that λij2\lambda_{ij}\geqslant 2. Then, RR still contains maximal subrings as in (4.2). But, JijJ_{ij} contains at least two distinct maximal subideals, so we have at least two choices for IijI_{ij}. Thus, RR contains at least 2qninj2q^{n_{i}n_{j}} maximal subrings of the form of (4.2). Appealing again to Lemma 3.5, we have

2qninjσ(R)<σ(Tij)qninj+11q1.2q^{n_{i}n_{j}}\leqslant\sigma(R)<\sigma(T_{ij})\leqslant\dfrac{q^{n_{i}n_{j}+1}-1}{q-1}.

But then, 2qninj(q1)<qninj+112q^{n_{i}n_{j}}(q-1)<q^{n_{i}n_{j}+1}-1, which is impossible. Hence, λij=1\lambda_{ij}=1 and JijJ_{ij} is simple.

(3) Suppose that both SiS_{i} and SjS_{j} are noncommutative, i.e., that ni2n_{i}\geqslant 2 and nj2n_{j}\geqslant 2. As in part (1), we obtain qninjσ(R)q^{n_{i}n_{j}}\leqslant\sigma(R) by considering maximal subrings of the form in (4.2). Without loss of generality, assume that ninjn_{i}\leqslant n_{j}. Since RR is σ\sigma-elementary, σ(R)<σ(R/J)σ(Si)\sigma(R)<\sigma(R/J)\leqslant\sigma(S_{i}). But,

σ(Si)=σ(Mni(qi))<qni2qninj\sigma(S_{i})=\sigma(M_{n_{i}}(q_{i}))<q^{n_{i}^{2}}\leqslant q^{n_{i}n_{j}}

a contradiction. Thus, at least one of SiS_{i} or SjS_{j} must be a field.

Remark 4.6.

Theorem 4.1 holds mutatis mutandis for unital maximal subrings and unital coverings of RR. This fact will be used later in Section 7 when we examine σu\sigma_{u}-elementary rings.

It remains possible that RR is σ\sigma-elementary and Jii{0}J_{ii}\neq\{0\} for some ii. Indeed, this case cannot be excluded entirely, because 𝔽q(+)𝔽q2\mathbb{F}_{q}(+)\mathbb{F}_{q}^{2} is σ\sigma-elementary and has a nonzero radical. However, we can restrict the possibilities for JiiJ_{ii} in this case as well. First, we show that if RR is σ\sigma-elementary and SiS_{i} is noncommutative, then Jii={0}J_{ii}=\{0\}.

Lemma 4.7.

Let 1iN1\leqslant i\leqslant N and let T=SiJiiT=S_{i}\oplus J_{ii}. Assume that ni2n_{i}\geqslant 2 and Jii{0}J_{ii}\neq\{0\}. Then, |JiiZ(T)|=qλii|J_{ii}\cap Z(T)|=q^{\lambda_{ii}}.

Proof.

For convenience, let Si=Mn(q)S_{i}=M_{n}(q), J=JiiJ=J_{ii}, and λ=λii\lambda=\lambda_{ii}. As in [28, Section 11.8], JJ decomposes as a direct sum k=1λXk\bigoplus_{k=1}^{\lambda}X_{k} of simple (Mn(q)M_{n}(q), Mn(q)M_{n}(q))-bimodules X1,,XλX_{1},\ldots,X_{\lambda}. Each XkX_{k} is equivalent to a simple Mn2(q)M_{n^{2}}(q) module—hence has order qn2q^{n^{2}}—and the isomorphism type of TT is determined entirely by λ\lambda. Thus, we can represent each XkX_{k} (additively) as a copy of Mn(q)M_{n}(q), and represent TT in block matrix form as

T={(Ax1xλ0A000A):AMn(q),xkXK for 1kλ}.T=\left\{\begin{pmatrix}A&x_{1}&\cdots&x_{\lambda}\\ 0&A&\cdots&0\\ \vdots&&\ddots&&\\ 0&\cdots&0&A\end{pmatrix}:A\in M_{n}(q),x_{k}\in X_{K}\text{ for }1\leqslant k\leqslant\lambda\right\}.

Via this representation, we see that xJZ(T)x\in J\cap Z(T) if and only if x=kxkx=\sum_{k}x_{k} and each xkx_{k} is an n×nn\times n scalar matrix. Thus, |JZ(T)|=qλ|J\cap Z(T)|=q^{\lambda}.∎

Proposition 4.8.

Let 1iN1\leqslant i\leqslant N and let T=SiJiiT=S_{i}\oplus J_{ii}. Assume that ni2n_{i}\geqslant 2.

  1. (1)

    TT is coverable, and σ(T)=σ(Mni(qi))\sigma(T)=\sigma(M_{n_{i}}(q_{i})).

  2. (2)

    If RR is σ\sigma-elementary, then Jii={0}J_{ii}=\{0\}.

Proof.

As in earlier proofs, we will drop some subscripts for readability. Let n=nin=n_{i}, q=qiq=q_{i}, and λ=λi\lambda=\lambda_{i}.

(1) The ring TT is coverable because it is noncommutative. For the covering number of TT, we use induction on λ\lambda. If λ=0\lambda=0, then T=Mn(q)T=M_{n}(q) and we are done. So, assume that λ1\lambda\geqslant 1 and that the covering number of T:=Mn(q)JT^{\prime}:=M_{n}(q)\oplus J^{\prime} is σ(Mn(q))\sigma(M_{n}(q)) whenever JJ^{\prime} is an (Mn(q)M_{n}(q), Mn(q)M_{n}(q))-bimodule of length λ1\lambda-1.

Consider M:=SiIM:=S_{i}\oplus I, where II is a maximal TT-subideal of JiiJ_{ii}. Note that II has length λ1\lambda-1, so σ(M)=σ(Mn(q))\sigma(M)=\sigma(M_{n}(q)) and |Jii:I|=qn2|J_{ii}:I|=q^{n^{2}}. On the one hand, II has an ideal complement in JiiJ_{ii}, so MM occurs as a residue ring of TT and σ(T)σ(M)\sigma(T)\leqslant\sigma(M). On the other hand, by Lemmas 3.2(3) and 4.7, the number of conjugates of MM in TT is

|Jii:(I+(JiiZ(T)))|=|Jii:I||IZ(T)||JiiZ(T)|=qn2qλ1qλ=qn21.|J_{ii}:(I+(J_{ii}\cap Z(T)))|=|J_{ii}:I|\cdot\frac{|I\cap Z(T)|}{|J_{ii}\cap Z(T)|}=q^{n^{2}}\cdot\frac{q^{\lambda-1}}{q^{\lambda}}=q^{n^{2}-1}.

By [33, Lemma 3.4], σ(M)qn2/2\sigma(M)\leqslant q^{n^{2}/2}, so any minimal cover of TT excludes some conjugate of MM. Hence, σ(M)σ(T)\sigma(M)\leqslant\sigma(T) by Lemma 2.2(1).

(2) Suppose that RR is σ\sigma-elementary, but Jii{0}J_{ii}\neq\{0\}. Let II be a maximal subideal of JiiJ_{ii}, let J^ii\widehat{J}_{ii} be an ideal complement for JiiJ_{ii} in JJ, and let S^i=kiSk\widehat{S}_{i}=\bigoplus_{k\neq i}S_{k}. Then, SiIS_{i}\oplus I is maximal in TT, so

M:=(SiS^i)(IJ^ii)M:=(S_{i}\oplus\widehat{S}_{i})\oplus(I\oplus\widehat{J}_{ii})

is a maximal subring of RR. As noted earlier, the number of conjugates of SiIS_{i}\oplus I in TT is qn21q^{n^{2}-1}. So, RR contains qn21q^{n^{2}-1} such maximal subrings MM. Since

σ(R)<σ(Si)=σ(Mn(q))<qn21,\sigma(R)<\sigma(S_{i})=\sigma(M_{n}(q))<q^{n^{2}-1},

every minimal cover of RR excludes at least one MM. By Lemma 2.2(1), σ(M)σ(R)\sigma(M)\leqslant\sigma(R). But, MM occurs as a residue ring of RR, so σ(R)<σ(M)\sigma(R)<\sigma(M), a contradiction. ∎

Next, we consider JkkJ_{kk} when SkS_{k} is a field. We will prove that for RR to be σ\sigma-elementary in this case, it is necessary that λkk2\lambda_{kk}\geqslant 2; and, if Jij{0}J_{ij}\neq\{0\} for some iji\neq j, then Jkk={0}J_{kk}=\{0\} whenever SkS_{k} is isomorphic to SiS_{i} or SjS_{j}.

Lemma 4.9.

Let 1kN1\leqslant k\leqslant N. Assume that SkS_{k} is a field and that JkkJ_{kk} is a simple (SkS_{k}, SkS_{k})-bimodule.

  1. (1)

    RR contains a unique maximal subring that does not contain JkkJ_{kk}.

  2. (2)

    σ(R)=σ(R/Jkk)\sigma(R)=\sigma(R/J_{kk}) and RR is not σ\sigma-elementary.

Proof.

(1) First, we note that JkkZ(R)J_{kk}\subseteq Z(R). Indeed, suppose that xJkkx\in J_{kk} and sSs\in S. Write s=i=1Nsis=\sum_{i=1}^{N}s_{i}, where each siSis_{i}\in S_{i}. Then, sx=skxsx=s_{k}x. But, SkJkkS_{k}\oplus J_{kk} is commutative, so skx=xsk=xss_{k}x=xs_{k}=xs.

Next, let J^kk\widehat{J}_{kk} be an ideal complement for JkkJ_{kk} in JJ, and let T=SJkkT=S\oplus J_{kk}. Then, TT is a maximal subring of RR because JkkJ_{kk} is simple. Let TT^{\prime} be any maximal subring of RR that does not contain JkkJ_{kk}. By Lemma 2.6, T=S1+yIT^{\prime}=S^{1+y}\oplus I for some yJy\in J and some maximal subideal II of JJ.

We claim that I=J^kkI=\widehat{J}_{kk}. To see this, apply the Peirce decomposition to II to get

I=(ekIek)((i,j)(k,k)eiIej)=(IJkk)(IJ^kk).I=(e_{k}Ie_{k})\oplus\Big{(}\bigoplus_{(i,j)\neq(k,k)}e_{i}Ie_{j}\Big{)}=(I\cap J_{kk})\oplus(I\cap\widehat{J}_{kk}).

Since JkkJ_{kk} is simple and JkkIJ_{kk}\not\subseteq I, IJkk={0}I\cap J_{kk}=\{0\}. Then, by maximality, I=J^kkI=\widehat{J}_{kk}.

Now, let αJkk\alpha\in J_{kk} and βJ^kk\beta\in\widehat{J}_{kk} be such that y=α+βy=\alpha+\beta. Since α\alpha is central, conjugation by 1+α1+\alpha fixes elements of SS pointwise. Thus,

S1+y=(S1+α)1+β=S1+βSJ^kkS^{1+y}=(S^{1+\alpha})^{1+\beta}=S^{1+\beta}\subseteq S\oplus\widehat{J}_{kk}

and so T=TT^{\prime}=T.

(2) As in part (1), let J^kk\widehat{J}_{kk} be the ideal complement for JkkJ_{kk} in JJ and let T:=SJ^kkT:=S\oplus\widehat{J}_{kk} be the unique maximal subring of RR that does not contain JkkJ_{kk}. Let 𝒞\mathcal{C} be a minimal cover of RR, and let A1,,AmA_{1},\ldots,A_{m} be all the maximal subrings in 𝒞\mathcal{C} that contain JkkJ_{kk}.

Note that R=TJkkR=T\oplus J_{kk}. Let aTa\in T, let xJkk{0}x\in J_{kk}\setminus\{0\}, and consider a+xa+x. Since SkTS_{k}\subseteq T and JkkJ_{kk} is simple, if xTx\in T then Skx=JkkTS_{k}\cdot x=J_{kk}\subseteq T, a contradiction. So, xTx\notin T, and hence a+xTa+x\notin T as well. Consequently, a+x=1mAa+x\in\bigcup_{\ell=1}^{m}A_{\ell}. This holds for all aTa\in T, so T+x=1mAT+x\subseteq\bigcup_{\ell=1}^{m}A_{\ell}. Recalling that JkkAJ_{kk}\subseteq A_{\ell} for each \ell, we see that {A1/Jkk,,Am/Jkk}\{A_{1}/J_{kk},\ldots,A_{m}/J_{kk}\} is a cover of R/JkkTR/J_{kk}\cong T. Thus, σ(R/Jkk)mσ(R)\sigma(R/J_{kk})\leqslant m\leqslant\sigma(R), and so in fact σ(R)=σ(R/Jkk)\sigma(R)=\sigma(R/J_{kk}). ∎

Remark 4.10.

We note that Lemma 4.9 holds mutatis mutandis for unital maximal subrings and unital coverings of RR. As with Theorem 4.1, this will become relevant in Section 7.

Proposition 4.11.

Let 1i,jN1\leqslant i,j\leqslant N such that iji\neq j. Assume that JijJ_{ij} is a simple (SiS_{i}, SjS_{j})-bimodule, SjS_{j} is a field, and RR is σ\sigma-elementary. Then, for all 1kN1\leqslant k\leqslant N, if SkSjS_{k}\cong S_{j}, then Jkk={0}J_{kk}=\{0\}.

Proof.

Assume 1kN1\leqslant k\leqslant N is such that SkSjS_{k}\cong S_{j}, and suppose that Jkk{0}J_{kk}\neq\{0\}. Since RR is σ\sigma-elementary, JkkJ_{kk} cannot be simple by Lemma 4.9. So, λkk2\lambda_{kk}\geqslant 2.

Now, SkJkkS_{k}\oplus J_{kk} is a commutative ring. Since λkk2\lambda_{kk}\geqslant 2 and Sk𝔽qjS_{k}\cong\mathbb{F}_{q_{j}}, the ring SkJkkS_{k}\oplus J_{kk} has 𝔽qj(+)𝔽qj2\mathbb{F}_{q_{j}}(+)\mathbb{F}_{q_{j}}^{2} as a residue ring. By [32, Theorem 4.8(2)],

σ(SkJkk)σ(𝔽qj(+)𝔽qj2)=qj+1.\sigma(S_{k}\oplus J_{kk})\leqslant\sigma(\mathbb{F}_{q_{j}}(+)\mathbb{F}_{q_{j}}^{2})=q_{j}+1.

Moreover, SkJkkS_{k}\oplus J_{kk} occurs as a residue ring of RR by Lemma 3.2(4), so σ(R)<qj+1\sigma(R)<q_{j}+1. However, letting q=qiqjq=q_{i}\otimes q_{j} and applying Lemma 3.6 yields

qj+1qni+1σ(R),q_{j}+1\leqslant q^{n_{i}}+1\leqslant\sigma(R),

which is a contradiction. Thus, Jkk={0}J_{kk}=\{0\}. ∎

Based on our work so far, it remains possible that Jij{0}J_{ij}\neq\{0\} and Jkk{0}J_{kk}\neq\{0\} for some field SkS_{k} that is isomorphic to neither SiS_{i} nor SjS_{j}. Ultimately, this case is ruled out by our classification of σ\sigma-elementary rings in Theorem 1.3. Rather than attack this case directly, however, we merely note that SkJkkS_{k}\oplus J_{kk} is always a commutative ring. The conclusion that Jkk={0}J_{kk}=\{0\} when Jij{0}J_{ij}\neq\{0\} will be reached later by referencing the classification of commutative σ\sigma-elementary rings from [32].

5. Restrictions on direct summands of SS

Continue to use the notations of Sections 3 and 4. At this point, we know that in a σ\sigma-elementary ring, the radical has a restricted structure. There is at most one pair (i,j)(i,j) with iji\neq j and Jij{0}J_{ij}\neq\{0\}, and if Jkk{0}J_{kk}\neq\{0\}, then SkS_{k} must be a field not isomorphic to SiS_{i} or SjS_{j}. We can also place restrictions on the direct summands of SS. In Proposition 5.3, we show that if SiS_{i} is noncommutative, then Sj≇SiS_{j}\not\cong S_{i} for all jij\neq i. A closely related result holds for a field SiS_{i} (Proposition 5.6): if Jij{0}J_{ij}\neq\{0\} with iji\neq j and SiS_{i} is a field, then Sk≇SiS_{k}\not\cong S_{i} for all ki,jk\neq i,j.

We begin by recalling some results from [27].

Lemma 5.1.

Let A=Mn(q)A=M_{n}(q).

  1. (1)

    [27, Proposition 4.2] Let ϕ\phi be an 𝔽p\mathbb{F}_{p}-automorphism of AA. Then, M:={a+ϕ(a):aA}M:=\{a+\phi(a):a\in A\} is a maximal subring of AAA\oplus A, and MAM\cong A.

  2. (2)

    [27, Lemma 5.5] If n2n\geqslant 2, then the number of 𝔽p\mathbb{F}_{p}-automorphisms of AA is greater than σ(A)\sigma(A).

Lemma 5.2.

Let 1iN1\leqslant i\leqslant N. Assume that Jij=Jji={0}J_{ij}=J_{ji}=\{0\} for all jij\neq i. Then, SiJiiS_{i}\oplus J_{ii} is a two-sided ideal of RR.

Proof.

Let R1:=SiJiiR_{1}:=S_{i}\oplus J_{ii} and

R2:=(jiSj)((j,k)(i,i)Jjk),R_{2}:=\Big{(}\bigoplus_{j\neq i}S_{j}\Big{)}\oplus\Big{(}\bigoplus_{(j,k)\neq(i,i)}J_{jk}\Big{)},

so that R=R1R2R=R_{1}\oplus R_{2}. By assumption, Jij={0}=JjiJ_{ij}=\{0\}=J_{ji} for all jij\neq i. So,

R2=(jiSj)(ji,kiJjk).R_{2}=\Big{(}\bigoplus_{j\neq i}S_{j}\Big{)}\oplus\Big{(}\bigoplus_{j\neq i,k\neq i}J_{jk}\Big{)}.

This means that r1r2=0=r2r1r_{1}r_{2}=0=r_{2}r_{1} for all r1R1r_{1}\in R_{1} and r2R2r_{2}\in R_{2}. It follows that R1R_{1} is a two-sided ideal of RR. ∎

Proposition 5.3.

Let 1i,jN1\leqslant i,j\leqslant N such that ni2n_{i}\geqslant 2. If RR is σ\sigma-elementary, then Si≇SjS_{i}\not\cong S_{j} for all jij\neq i.

Proof.

Assume that RR is σ\sigma-elementary, but there exists jij\neq i such that SjSiS_{j}\cong S_{i}. By Theorem 4.1, there is at most one pair (k,)(k,\ell) with kk\neq\ell and Jk{0}J_{k\ell}\neq\{0\}, and we must have Jij=Jji={0}J_{ij}=J_{ji}=\{0\} because SiS_{i} and SjS_{j} are noncommutative. Also, Jii=Jjj={0}J_{ii}=J_{jj}=\{0\} by Proposition 4.8. It follows that either Jik=Jki={0}J_{ik}=J_{ki}=\{0\} for all 1kN1\leqslant k\leqslant N, or Jjk=Jkj={0}J_{jk}=J_{kj}=\{0\} for all 1kN1\leqslant k\leqslant N. Assume without loss of generality that the latter condition holds. Then, SjS_{j} is a two-sided ideal of RR by Lemma 5.2.

For any 𝔽p\mathbb{F}_{p}-algebra isomorphism ϕ:SiSj\phi:S_{i}\to S_{j}, let MϕM_{\phi} be the maximal subring Mϕ:={a+ϕ(a):aSi}M_{\phi}:=\{a+\phi(a):a\in S_{i}\} of SiSjS_{i}\oplus S_{j}. Note that Mϕ1=Mϕ2M_{\phi_{1}}=M_{\phi_{2}} if and only if ϕ1=ϕ2\phi_{1}=\phi_{2}. Let

T:=(ki,kjSk)MϕJ,T:=\Big{(}\bigoplus_{k\neq i,k\neq j}S_{k}\Big{)}\oplus M_{\phi}\oplus J,

which is maximal subring of RR. Furthermore, since MϕSiSjM_{\phi}\cong S_{i}\cong S_{j}, we have

T(kjSk)JR/Sj,T\cong\Big{(}\bigoplus_{k\neq j}S_{k}\Big{)}\oplus J\cong R/S_{j},

and so σ(R)<σ(R/Sj)=σ(T)\sigma(R)<\sigma(R/S_{j})=\sigma(T).

Now, let mm be the number of maximal subrings of RR isomorphic to TT. Then, mm is greater than or equal to the number of choices for MϕM_{\phi}, which by Lemma 5.1(2) is greater than σ(Mni(qi))=σ(Si)\sigma(M_{n_{i}}(q_{i}))=\sigma(S_{i}). Since RR has a residue ring isomorphic to SiS_{i}, we have σ(R)<σ(Si)<m\sigma(R)<\sigma(S_{i})<m. Hence, any minimal cover of RR omits some maximal subring TT^{\prime} isomorphic to TT. Applying Lemma 2.2(1), we have σ(T)=σ(T)σ(R)\sigma(T)=\sigma(T^{\prime})\leqslant\sigma(R), which is a contradiction. ∎

Ruling out fields isomorphic to SjS_{j} is more difficult. By [35, Theorem 3.5], T==1t𝔽qT=\bigoplus_{\ell=1}^{t}\mathbb{F}_{q} is coverable if and only if tτ(q)t\geqslant\tau(q), where τ(q)\tau(q) is as defined in the introduction. When TT is not coverable, it is generated (as a ring) by a single element α==1tα\alpha=\sum_{\ell=1}^{t}\alpha_{\ell}. By [35, Proposition 3.4], α\alpha generates TT if and only if each α\alpha_{\ell} is a primitive element for 𝔽q\mathbb{F}_{q} over 𝔽p\mathbb{F}_{p} (if q=pq=p, then any nonzero element of 𝔽p\mathbb{F}_{p} is primitive) and each α\alpha_{\ell} has a distinct minimal polynomial over 𝔽p\mathbb{F}_{p}.

We also require knowledge about the maximal subrings of SS. These subrings were classified in [27].

Definition 5.4.

[27, Definition 4.3] Let S=i=1NSiS=\bigoplus_{i=1}^{N}S_{i}, where each SiS_{i} is a finite simple ring of characteristic pp. Let MM be a maximal subring of SS. We say that MM is of Type Π1\Pi_{1} if there exists an index ii and a maximal subring MiM_{i} of RiR_{i} such that

M=S1Si1MiSi+1SN.M=S_{1}\oplus\cdots\oplus S_{i-1}\oplus M_{i}\oplus S_{i+1}\oplus\cdots\oplus S_{N}.

We say that MM is of Type Π2\Pi_{2} if there exist indices i<ji<j and an 𝔽p\mathbb{F}_{p}-algebra isomorphism ϕ:SiSj\phi:S_{i}\to S_{j} such that

M={k=1NskS:skSk for each 1kN and sj=ϕ(si)};M=\Big{\{}\sum_{k=1}^{N}s_{k}\in S:s_{k}\in S_{k}\text{ for each }1\leqslant k\leqslant N\text{ and }s_{j}=\phi(s_{i})\Big{\}};

here, there are no restrictions on the summands sks_{k} for ki,jk\neq i,j. In a Type Π2\Pi_{2} maximal subring, we say that SiS_{i} and SjS_{j} are linked by ϕ\phi.

Lemma 5.5.

[27, Theorem 4.5] Any maximal subring of SS is Type Π1\Pi_{1} or Type Π2\Pi_{2}.

Proposition 5.6.

Let 1i,jN1\leqslant i,j\leqslant N with iji\neq j. Assume that Jij{0}J_{ij}\neq\{0\} and SjS_{j} is a field. If RR is σ\sigma-elementary, then Sk≇SjS_{k}\not\cong S_{j} for all ki,jk\neq i,j.

Proof.

Assume that RR is σ\sigma-elementary. By Theorem 4.1, JijJ_{ij} is a simple (SiS_{i}, SjS_{j})-bimodule. Let 𝒯:={1kN:ki,kj,SkSj}\mathscr{T}:=\{1\leqslant k\leqslant N:k\neq i,k\neq j,S_{k}\cong S_{j}\}. Suppose that 𝒯\mathscr{T}\neq\varnothing, let t=|𝒯|t=|\mathscr{T}|, and let T=k𝒯SkT=\bigoplus_{k\in\mathscr{T}}S_{k}. From Theorem 4.1, we know that Jk=Jk={0}J_{k\ell}=J_{\ell k}=\{0\} whenever k𝒯k\in\mathscr{T} and k\ell\neq k. Moreover, by Proposition 4.11, Jkk={0}J_{kk}=\{0\} for all k𝒯k\in\mathscr{T}. Hence, by Lemma 5.2, each SkS_{k}—and therefore TT—is a two-sided ideal of RR.

Let qj=pdjq_{j}=p^{d_{j}} and q=qiqjq=q_{i}\otimes q_{j}. If ni=1n_{i}=1, then R=SiSjJijR=S_{i}\oplus S_{j}\oplus J_{ij} by Lemma 3.6, and there is nothing to prove because 𝒯=\mathscr{T}=\varnothing. So, we may assume that ni2n_{i}\geqslant 2.

We break the proof into two cases, depending on whether t<τ(qj)1t<\tau(q_{j})-1 or tτ(qj)1t\geqslant\tau(q_{j})-1. The latter case is easier, so we will deal with that first.

When tτ(qj)1t\geqslant\tau(q_{j})-1, SjTi=1t+1𝔽qjS_{j}\oplus T\cong\bigoplus_{i=1}^{t+1}\mathbb{F}_{q_{j}} is coverable, and σ(R)<σ(SjT)\sigma(R)<\sigma(S_{j}\oplus T) since SjTS_{j}\oplus T occurs as a residue ring of RR. By Lemma 3.7,

(5.7) σ(R)<{12(p2+p),dj=1qj2/(2dj),dj2.\sigma(R)<\begin{cases}\tfrac{1}{2}(p^{2}+p),&d_{j}=1\\ q_{j}^{2}/(2d_{j}),&d_{j}\geqslant 2\end{cases}.

However, by Lemma 3.6, qj2+1qni+1σ(R)q_{j}^{2}+1\leqslant q^{n_{i}}+1\leqslant\sigma(R). This is not compatible with (5.7), since neither p2+1<12(p2+p)p^{2}+1<\tfrac{1}{2}(p^{2}+p) nor qj2+1<qj2/(2dj)q_{j}^{2}+1<q_{j}^{2}/(2d_{j}) is a valid inequality.

For the remainder of the proof, assume that t<τ(qj)1t<\tau(q_{j})-1. Define T^\widehat{T} to be T^:=𝒯S\widehat{T}:=\bigoplus_{\ell\notin\mathscr{T}}S_{\ell}. Then, R=(T^J)TR=(\widehat{T}\oplus J)\oplus T.

We will sort the maximal subrings of RR into three classes 𝒞1\mathcal{C}_{1}, 𝒞2\mathcal{C}_{2}, and 𝒞3\mathcal{C}_{3}. Let 𝒞1\mathcal{C}_{1} be the collection of all maximal subrings of RR that contain TT. Note that these maximal subrings are all inverse images of maximal subrings of R/TR/T.

Next, suppose MRM\subseteq R is maximal with TMT\not\subseteq M. If JMJ\not\subseteq M, then M=SIM=S^{\prime}\oplus I for some S𝒮(R)S^{\prime}\in\mathscr{S}(R) and some maximal subideal IJI\subseteq J. However, as noted earlier, Jk=Jk={0}J_{k\ell}=J_{\ell k}=\{0\} for all 1N1\leqslant\ell\leqslant N and all k𝒯k\in\mathscr{T}. So, the action of TT on JJ is trivial. It follows that TST\subseteq S^{\prime} for all S𝒮(R)S^{\prime}\in\mathscr{S}(R), and hence JJ must be contained in MM.

Since JMJ\subseteq M and TMT\not\subseteq M, MM is the inverse image of a maximal subring of R/JR/J that does not contain TT. In light of Lemma 5.5, we can partition such subrings into two classes. Use a bar to denote passage from RR to R/JR/J. Let 𝒞2\mathcal{C}_{2} be the set of all subrings of the form T^JM\widehat{T}\oplus J\oplus M^{\prime}, where MM^{\prime} is maximal in TT. Subrings in 𝒞2\mathcal{C}_{2} are those where Sj¯\overline{S_{j}} is not linked to any Sk¯\overline{S_{k}}, k𝒯k\in\mathscr{T}. Finally, let 𝒞3\mathcal{C}_{3} be the collection of maximal subrings where Sj¯\overline{S_{j}} is linked to some Sk¯\overline{S_{k}} with k𝒯k\in\mathscr{T}.

Now, since t<τ(qj)t<\tau(q_{j}), TT is not coverable. Next, we will show that for all aT^Ja\in\widehat{T}\oplus J, we can find αT\alpha\in T such that α\alpha generates TT (as a ring), and a+αa+\alpha is not contained in a maximal subring in either 𝒞2\mathcal{C}_{2} or 𝒞3\mathcal{C}_{3}. This is easy to see with regard to 𝒞2\mathcal{C}_{2}: for any aa and α\alpha as above, if a+αM𝒞2a+\alpha\in M\in\mathcal{C}_{2}, then αM\alpha\in M^{\prime} for some maximal subring MM^{\prime} of TT. But, this is impossible because α\alpha generates TT.

For 𝒞3\mathcal{C}_{3}, the argument is more delicate. Let S^j\widehat{S}_{j} be an additive complement to SjS_{j} in T^\widehat{T}. Write aa as a=s+ba=s+b, where sSjs\in S_{j} and bS^jJb\in\widehat{S}_{j}\oplus J. Since t<τ(qj)1t<\tau(q_{j})-1, t+1t+1 is less than or equal to the number of irreducible polynomials in 𝔽p[x]\mathbb{F}_{p}[x] of degree djd_{j}. So, for each k𝒯k\in\mathscr{T}, we can find αkSk\alpha_{k}\in S_{k} such that αk\alpha_{k} generates SkS_{k}; each αk\alpha_{k} has a distinct minimal polynomial over 𝔽p\mathbb{F}_{p}; and no αk\alpha_{k} has the same minimal polynomial as ss. Let α=k𝒯αk\alpha=\sum_{k\in\mathscr{T}}\alpha_{k}. Then, for every kk and every automorphism ϕ\phi of 𝔽qj\mathbb{F}_{q_{j}}, ϕ(s)αk\phi(s)\neq\alpha_{k}. We conclude that a+αa+\alpha cannot lie in any subring where Sj¯\overline{S_{j}} is linked to some Sk¯\overline{S_{k}}.

We have shown that for every aT^Ja\in\widehat{T}\oplus J, there exists αT\alpha\in T such that a+αa+\alpha lies in a maximal subring in class 𝒞1\mathcal{C}_{1}. Now, consider a minimal cover 𝒞\mathcal{C} of RR. Let 𝒜=𝒞𝒞1\mathcal{A}=\mathcal{C}\cap\mathcal{C}_{1}. By our work above, 𝒜\mathcal{A}\neq\varnothing, because elements of the form a+αa+\alpha can only lie in maximal subrings in 𝒞1\mathcal{C}_{1}. Furthermore, the subrings in 𝒜\mathcal{A} form a cover of R/TR/T, so σ(R/T)|𝒜|σ(R)\sigma(R/T)\leqslant|\mathcal{A}|\leqslant\sigma(R). This contradicts the fact that RR is σ\sigma-elementary, and we are done.∎

6. Proof of Theorem 1.3

We now have all the ingredients necessary to prove Theorem 1.3. Let RR be a σ\sigma-elementary ring with unity. The idea of the proof is to write RR as a direct sum of subrings R=R1R2R3R=R_{1}\oplus R_{2}\oplus R_{3} in such a way that all maximal subrings of RR respect the decomposition. That is, any maximal subring MM of RR is equal to one of

(6.1) M1R2R3,R1M2R3,orR1R2M3,M_{1}\oplus R_{2}\oplus R_{3},\;R_{1}\oplus M_{2}\oplus R_{3},\;\text{or}\;R_{1}\oplus R_{2}\oplus M_{3},

where MiM_{i} is maximal in RiR_{i}. Then, σ(R)=min{σ(R1),σ(R2),σ(R3)}\sigma(R)=\min\{\sigma(R_{1}),\sigma(R_{2}),\sigma(R_{3})\} by Lemma 2.3, and the problem is reduced to a small number of possible cases.

Proposition 6.2.

Assume RR is a finite unital ring of characteristic pp containing subrings R1R_{1}, R2R_{2}, and R3R_{3} that satisfy all of the following properties:

  • R=R1R2R3R=R_{1}\oplus R_{2}\oplus R_{3}.

  • R1A(n,q1,q2)R_{1}\cong A(n,q_{1},q_{2}) for some n2n\geqslant 2.

  • R2jMnj(qj)R_{2}\cong\bigoplus_{j\in\mathcal{I}}M_{n_{j}}(q_{j}), where \mathcal{I} is some finite index set, nj2n_{j}\geqslant 2 for each jj, and Mnj(qj)≇Mn(q1)M_{n_{j}}(q_{j})\not\cong M_{n}(q_{1}) for all jj.

  • R3R_{3} is commutative, and if a field FF occurs as a direct summand of R3/𝒥(R3)R_{3}/\mathscr{J}(R_{3}), then F≇𝔽q2F\not\cong\mathbb{F}_{q_{2}}.

Then, σ(R)=min{σ(R1),σ(R2),σ(R3)}\sigma(R)=\min\{\sigma(R_{1}),\sigma(R_{2}),\sigma(R_{3})\}.

Proof.

It suffices to show that every maximal subring of RR decomposes as in (6.1). Then, Lemma 2.3 can be used to reach the desired conclusion.

Let J=𝒥(R)J=\mathscr{J}(R). For i=1,2,3i=1,2,3, let Ji=𝒥(Ri)J_{i}=\mathscr{J}(R_{i}), and fix a semisimple complement SiS_{i} for JiJ_{i} in RiR_{i}. Note that J2={0}J_{2}=\{0\} and S2=R2S_{2}=R_{2} because R2R_{2} is semisimple. So, J=J1J3J=J_{1}\oplus J_{3} and S:=S1R2S3S:=S_{1}\oplus R_{2}\oplus S_{3} is a semisimple complement to JJ in RR.

Let MM be a maximal subring of RR. Assume first that JMJ\not\subseteq M. By Lemma 2.6, M=SIM=S^{\prime}\oplus I, where SS^{\prime} is a conjugate of SS and II is a maximal subideal of JJ. Any such II decomposes as either I=I1J3I=I_{1}\oplus J_{3} or I=J1I3I=J_{1}\oplus I_{3}, where I1I_{1} (respectively, I3I_{3}) is maximal in J1J_{1} (respectively, J3J_{3}). Moreover, the conjugates of SS have a similar form. Indeed, S1S2S_{1}\oplus S_{2} has a trivial action on J3J_{3}, and likewise for the action of S2S3S_{2}\oplus S_{3} on J1J_{1}. So, given x=x1+x3Jx=x_{1}+x_{3}\in J with x1J1x_{1}\in J_{1} and x3J3x_{3}\in J_{3}, we have S1+x=S11+x1R2S31+x3S^{1+x}=S_{1}^{1+x_{1}}\oplus R_{2}\oplus S_{3}^{1+x_{3}}.

With these observations in mind, we see that if I=I1J3I=I_{1}\oplus J_{3}, then

M=SI=(S1I1)R2(S3J3)\displaystyle M=S^{\prime}\oplus I=(S_{1}^{\prime}\oplus I_{1})\oplus R_{2}\oplus(S_{3}^{\prime}\oplus J_{3})

for some conjugate S1S_{1}^{\prime} of S1S_{1} and some conjugate S3S_{3}^{\prime} of S3S_{3}. In this case, S1I1S_{1}^{\prime}\oplus I_{1} is maximal in R1R_{1}, and S3J3=R3S_{3}^{\prime}\oplus J_{3}=R_{3}. Thus, MM has the form of (6.1). The same steps show that this also holds when I=J1I3I=J_{1}\oplus I_{3}.

Now, assume that JMJ\subseteq M. Then, M=TJM=T\oplus J for some maximal subring TT of SS. By Lemma 5.5, TT is either of Type Π1\Pi_{1} or Type Π2\Pi_{2}. If TT is of Type Π1\Pi_{1}, then TT has one of the forms

T1S2S3,S1T2S3, or S1S2T3,T_{1}\oplus S_{2}\oplus S_{3},\;S_{1}\oplus T_{2}\oplus S_{3},\;\text{ or }\;S_{1}\oplus S_{2}\oplus T_{3},

where TiT_{i} is maximal in SiS_{i} for i=1,2,3i=1,2,3, and MM decomposes as in (6.1). On the other hand, if TT is of Type Π2\Pi_{2}, then (modulo JJ) two simple summands AA and AA^{\prime} of SS are linked by an 𝔽p\mathbb{F}_{p}-algebra isomorphism. If AA is a field, then both AA and AA^{\prime} are subrings of S3S_{3}, because no simple summand of S3S_{3} is isomorphic to 𝔽q2\mathbb{F}_{q_{2}}. In this case, T=S1S2T3T=S_{1}\oplus S_{2}\oplus T_{3}. Similarly, if AA is noncommutative, then both AA and AA^{\prime} are subrings of S2S_{2}, because no simple summand of S2S_{2} is isomorphic to Mn(q1)M_{n}(q_{1}). Hence, T=S1T2S3T=S_{1}\oplus T_{2}\oplus S_{3}. Thus, in all cases MM has the form of (6.1), and we are done. ∎

Proof of Theorem 1.3.

We may assume that RR is finite, has characteristic pp, and J:=𝒥(R)J:=\mathscr{J}(R) satisfies J2={0}J^{2}=\{0\}. We will use the notation established at the beginning of Section 3. By Theorem 4.1, there is at most pair (i,j)(i,j) such that 1i,jN1\leqslant i,j\leqslant N, iji\neq j, and Jij{0}J_{ij}\neq\{0\}. If such a pair exists, then JijJ_{ij} is simple and one of SiS_{i} or SjS_{j} is a field. Without loss of generality, assume that nj=1n_{j}=1. Then, SiSjJijA(ni,qi,qj)S_{i}\oplus S_{j}\oplus J_{ij}\cong A(n_{i},q_{i},q_{j}). If ni=1n_{i}=1, then R=SiSjJijR=S_{i}\oplus S_{j}\oplus J_{ij} by Lemma 3.6 and we are done. So, assume that ni2n_{i}\geqslant 2.

If 1kN1\leqslant k\leqslant N is such that SkS_{k} is noncommutative, then Jkk={0}J_{kk}=\{0\} by Proposition 4.8, and by Proposition 5.3, S≇SkS_{\ell}\not\cong S_{k} for all k\ell\neq k. In particular, Jii={0}J_{ii}=\{0\} and Sk≇SiS_{k}\not\cong S_{i} when kik\neq i. Likewise, Jjj={0}J_{jj}=\{0\} by Proposition 4.11, and for all 1kN1\leqslant k\leqslant N such that kjk\neq j, we have Sk≇SjS_{k}\not\cong S_{j}.

Form the following three subrings of RR (these subrings may be equal to {0}\{0\}). Let R1:=SiSjJijR_{1}:=S_{i}\oplus S_{j}\oplus J_{ij}. Let :={1kN:ki,kj,Sk is noncommutative}\mathcal{I}:=\{1\leqslant k\leqslant N:k\neq i,k\neq j,S_{k}\text{ is noncommutative}\} and let R2:=kSkR_{2}:=\bigoplus_{k\in\mathcal{I}}S_{k}. Finally, let :={1kN:ki,kj,k}\mathcal{I}^{\prime}:=\{1\leqslant k\leqslant N:k\neq i,k\neq j,k\notin\mathcal{I}\} and let R3:=k(SkJkk)R_{3}:=\bigoplus_{k\in\mathcal{I}^{\prime}}(S_{k}\oplus J_{kk}). Then, R=R1R2R3R=R_{1}\oplus R_{2}\oplus R_{3}, and the subrings satisfy the hypotheses of Proposition 6.2. Thus, σ(R)=min{σ(R1),σ(R2),σ(R3)}\sigma(R)=\min\{\sigma(R_{1}),\sigma(R_{2}),\sigma(R_{3})\}. But, RR is σ\sigma-elementary, so in fact R=RR=R_{\ell} for some 131\leqslant\ell\leqslant 3, and the other two subrings must be {0}\{0\}. If R=R1R=R_{1}, then RA(ni,qi,qj)R\cong A(n_{i},q_{i},q_{j}). If R=R2R=R_{2}, then RR is noncommutative and semisimple. Hence, RMn(q)R\cong M_{n}(q) for some n2n\geqslant 2 by [27, Proposition 5.4, Theorem 5.11]. Finally, if R=R3R=R_{3}, then RR is commutative. We may then apply [32, Theorem 4.8] to conclude that either Ri=1τ(q)𝔽qR\cong\bigoplus_{i=1}^{\tau(q)}\mathbb{F}_{q}, or R𝔽q(+)𝔽q2R\cong\mathbb{F}_{q}(+)\mathbb{F}_{q}^{2}, depending on whether RR is semisimple or not. ∎

7. Rings without unity and covers by unital subrings

Theorem 1.3 provides a complete classification of σ\sigma-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 RR must contain 1R1_{R}? 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 σu\sigma_{u} and σu\sigma_{u}-elementary given in Definition 1.4. It is clear that parts (3) and (4) of Lemma 2.1 carry over to coverings of RR by unital subrings: for any two-sided ideal of RR, we have σu(R)σu(R/I)\sigma_{u}(R)\leqslant\sigma_{u}(R/I); and when RR 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 RR is finite and has prime power order. By results of Neumann [26, Lemma 4.1,4.4] and Lewin [22, Lemma 1], RR contains a two-sided ideal II of finite index such that σ(R)=σ(R/I)\sigma(R)=\sigma(R/I). Hence, we may assume that RR is finite. Next, the Chinese Remainder Theorem still applies to finite rings without identity, so Ri=1tRiR\cong\bigoplus_{i=1}^{t}R_{i}, where each RiR_{i} is a ring and |Ri|=pini|R_{i}|=p_{i}^{n_{i}} for some distinct primes p1,,ptp_{1},\ldots,p_{t}. Lemma 2.3 holds for rings without unity, so σ(R)=min1it{σ(Ri)}\sigma(R)=\min_{1\leqslant i\leqslant t}\{\sigma(R_{i})\}. Thus, we may assume that |R|=pn|R|=p^{n} for some prime pp and some n1n\geqslant 1.

Now, as in [32, Lemma 3.2, Proposition 3.3], one may prove that pRpR is contained in every maximal subring of RR. It follows that σ(R)=σ(R/pR)\sigma(R)=\sigma(R/pR), and hence we may assume that pR={0}pR=\{0\}. From here, we will embed RR in a unital ring RR^{\prime} that has characteristic pp.

Let R:=𝔽p×RR^{\prime}:=\mathbb{F}_{p}\times R with multiplication given by the rule:

(n1,r1)(n2,r2):=(n1n2,n1r2+n2r1+r1r2).(n_{1},r_{1})(n_{2},r_{2}):=(n_{1}n_{2},n_{1}r_{2}+n_{2}r_{1}+r_{1}r_{2}).

Notice that, for any (n,r)R(n,r)\in R^{\prime}, (1,0)(n,r)=(n,r)(1,0)=(n,r)(1,0)(n,r)=(n,r)(1,0)=(n,r), so RR^{\prime} has a multiplicative identity. We claim that σu(R)=σ(R)\sigma_{u}(R^{\prime})=\sigma(R).

Indeed, let 𝒞\mathcal{C} be a minimal cover of RR by subrings. Given S𝒞S\in\mathcal{C}, define 𝔽p×S:={(n,s)R:sS}\mathbb{F}_{p}\times S:=\{(n,s)\in R^{\prime}:s\in S\}, which is a unital subring of RR^{\prime}, and is proper in RR^{\prime} because SRS\subsetneq R. Let 𝒞:={𝔽p×S:S𝒞}\mathcal{C}^{\prime}:=\{\mathbb{F}_{p}\times S:S\in\mathcal{C}\}. If (n,r)R(n,r)\in R^{\prime}, then rSr\in S for some S𝒞S\in\mathcal{C}, so (n,r)𝔽p×S𝒞(n,r)\in\mathbb{F}_{p}\times S\in\mathcal{C}^{\prime}. Hence, 𝒞\mathcal{C}^{\prime} is a cover of RR^{\prime} by unital subrings, and so σu(R)|𝒞|=|𝒞|=σ(R).\sigma_{u}(R^{\prime})\leqslant|\mathcal{C}^{\prime}|=|\mathcal{C}|=\sigma(R).

Next, let 𝒰\mathcal{U}^{\prime} be a minimal cover of RR^{\prime} by unital subrings, which we now know must exist. Let ρ:RR\rho:R^{\prime}\to R be the projection map defined by ρ((n,r)):=r\rho((n,r)):=r. The map ρ\rho is not multiplicative, but ρ(S)\rho(S^{\prime}) is a proper subring of RR for any S𝒰S^{\prime}\in\mathcal{U}^{\prime}. To see this, let r1,r2ρ(S)r_{1},r_{2}\in\rho(S^{\prime}). Then, (n1,r1),(n2,r2)S(n_{1},r_{1}),(n_{2},r_{2})\in S^{\prime} for some n1,n2𝔽pn_{1},n_{2}\in\mathbb{F}_{p}. Because (1,0)S(1,0)\in S^{\prime}, both (n1,0)(n_{1},0) and (n2,0)(n_{2},0) are in SS^{\prime}. Thus, SS^{\prime} contains both (0,r1)(0,r_{1}) and (0,r2)(0,r_{2}), and hence r1+r2,r1r2ρ(S)r_{1}+r_{2},r_{1}r_{2}\in\rho(S^{\prime}). Moreover, ρ(S)R\rho(S^{\prime})\neq R, because if not, then {0}×RS\{0\}\times R\subseteq S^{\prime} and (1,0)S(1,0)\in S^{\prime}, which implies that S=RS^{\prime}=R^{\prime}.

Finally, given rRr\in R, there exists S𝒰S^{\prime}\in\mathcal{U}^{\prime} such that (0,r)S(0,r)\in S^{\prime}. Hence, rϕ(S)r\in\phi(S^{\prime}) and {ϕ(S):S𝒰}\{\phi(S^{\prime}):S^{\prime}\in\mathcal{U}^{\prime}\} is a cover of RR of size at most |𝒰||\mathcal{U}^{\prime}|. Consequently, σ(R)|𝒰|=σu(R)\sigma(R)\leqslant|\mathcal{U}^{\prime}|=\sigma_{u}(R), 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 RR, for which we may consider either covers by subrings (which need not contain 1R1_{R}) or covers by unital subrings. These problems can in turn be solved by classifying σ\sigma-elementary and σu\sigma_{u}-elementary rings. The classification of σ\sigma-elementary rings has been done in Theorem 1.3. We now proceed to prove Theorem 1.6, which classifies σu\sigma_{u}-elementary rings. As we shall see, the class of σu\sigma_{u}-elementary rings is largely the same as the class of σ\sigma-elementary rings, with the greatest discrepancy occurring in the commutative semisimple case.

Lemma 7.1.

Let RR be a ring with unity. Then, σ(R)σu(R)\sigma(R)\leqslant\sigma_{u}(R), with equality if and only if the subrings in some minimal cover of RR each contain 1R1_{R}.

Proof.

Note that every unital subring is a subring (but not conversely). ∎

Lemma 7.2.

  1. (1)

    For any prime pp, i=1p𝔽p\bigoplus_{i=1}^{p}\mathbb{F}_{p} is σ\sigma-elementary, but is not coverable by unital subrings.

  2. (2)

    For any prime pp, i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p} is σu\sigma_{u}-elementary, but is not σ\sigma-elementary. Furthermore, σu(i=1p+1𝔽p)=σ(i=1p+1𝔽p)=p+(p2)\sigma_{u}\big{(}\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}\big{)}=\sigma\big{(}\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}\big{)}=p+\binom{p}{2}.

  3. (3)

    The ring A(1,2,2)A(1,2,2) is σu\sigma_{u}-elementary, but is not σ\sigma-elementary. Furthermore, σu(A(1,2,2))=σ(A(1,2,2))=3\sigma_{u}(A(1,2,2))=\sigma(A(1,2,2))=3.

Proof.

(1) Let R=i=1p𝔽pR=\bigoplus_{i=1}^{p}\mathbb{F}_{p}. By [35, Theorem 3.5], RR is coverable, but i=1p1𝔽p\bigoplus_{i=1}^{p-1}\mathbb{F}_{p} is not, so RR is σ\sigma-elementary. As shown in [35, Theorems 5.3], a minimal cover of RR requires every maximal subring of RR. Since {0}\{0\} is the only maximal subring of 𝔽p\mathbb{F}_{p}, the Type Π1\Pi_{1} (see Definition 5.4) maximal subrings of RR do not contain 1R1_{R}. Thus, RR is not coverable by unital subrings.

(2) Let Ri𝔽pR_{i}\cong\mathbb{F}_{p} for 1ip+11\leqslant i\leqslant p+1, and let R=i=1p+1RiR=\bigoplus_{i=1}^{p+1}R_{i}. Given a=i=1p+1aiRa=\sum_{i=1}^{p+1}a_{i}\in R with each aiRia_{i}\in R_{i}, we must have ai=aja_{i}=a_{j} for some iji\neq j. Thus, aa lies in the Type Π2\Pi_{2} maximal subring of RR where RiR_{i} is linked to RjR_{j}. There are (p+12)=p+(p2)\binom{p+1}{2}=p+\binom{p}{2} such maximal subrings, and each one contains 1R1_{R}. Thus, σu(R)p+(p2)\sigma_{u}(R)\leqslant p+\binom{p}{2}; but, p+(p2)=σ(R)σu(R)p+\binom{p}{2}=\sigma(R)\leqslant\sigma_{u}(R) by Lemma 7.1, so σu(R)=p+(p2)\sigma_{u}(R)=p+\binom{p}{2}. By part (1), RR is σu\sigma_{u}-elementary. However, RR has i=1p𝔽p\bigoplus_{i=1}^{p}\mathbb{F}_{p} as a residue ring, and thus is not σ\sigma-elementary.

(3) Let R=A(1,2,2)R=A(1,2,2), which is equal to the ring of upper triangular matrices over 𝔽2\mathbb{F}_{2}. The three unital subrings S1:={(a00b):a,b𝔽2}S_{1}:=\left\{\left(\begin{smallmatrix}a&0\\ 0&b\end{smallmatrix}\right):a,b\in\mathbb{F}_{2}\ \right\}, S2:={(ab0a):a,b𝔽2}S_{2}:=\left\{\left(\begin{smallmatrix}a&b\\ 0&a\end{smallmatrix}\right):a,b\in\mathbb{F}_{2}\ \right\}, and S3:={(aa+b0b):a,b𝔽2}S_{3}:=\left\{\left(\begin{smallmatrix}a&a+b\\ 0&b\end{smallmatrix}\right):a,b\in\mathbb{F}_{2}\ \right\} form a cover of RR, so σu(R)=3\sigma_{u}(R)=3. The nonzero, proper residue rings of RR are isomorphic to 𝔽2𝔽2\mathbb{F}_{2}\oplus\mathbb{F}_{2} and 𝔽2\mathbb{F}_{2}, neither of which has a unital cover. Hence, RR is σu\sigma_{u}-elementary. Finally, since σ(𝔽2𝔽2)=3\sigma(\mathbb{F}_{2}\oplus\mathbb{F}_{2})=3, RR is not σ\sigma-elementary. ∎

Proposition 7.3.

Let RR be a σ\sigma-elementary ring with unity. Then, RR is σu\sigma_{u}-elementary if and only if R≇i=1p𝔽pR\not\cong\bigoplus_{i=1}^{p}\mathbb{F}_{p}. Moreover, when RR is σu\sigma_{u}-elementary, we have σu(R)=σ(R)\sigma_{u}(R)=\sigma(R).

Proof.

We show that, with the exception of i=1p𝔽p\bigoplus_{i=1}^{p}\mathbb{F}_{p}, all of the σ\sigma-elementary rings listed in Theorem 1.3 admit minimal covers by unital subrings. Indeed, if qpq\neq p, then every maximal subring of i=1τ(q)𝔽q\bigoplus_{i=1}^{\tau(q)}\mathbb{F}_{q} contains the identity of the ring [35, Theorem 4.8]. For any qq, a cover of 𝔽q(+)𝔽q2\mathbb{F}_{q}(+)\mathbb{F}_{q}^{2} by unital maximal subrings is given in [35, Example 6.1]. For n2n\geqslant 2, maximal subrings of Mn(q)M_{n}(q) 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 A(n,q1,q2)A(n,q_{1},q_{2}) is σ\sigma-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 RR is σu\sigma_{u}-elementary but not σ\sigma-elementary, then either Ri=1p+1𝔽pR\cong\bigoplus_{i=1}^{p+1}\mathbb{F}_{p} or RA(1,2,2)R\cong A(1,2,2). 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 pp with 2-nilpotent radicals. For the remainder of this section, we will assume the following:

Notation 7.4.

Let RR be a unital ring with characteristic pp, and let J:=𝒥(R)J:=\mathscr{J}(R) be the Jacobson radical of RR. Assume that J2={0}J^{2}=\{0\}. Let 𝒮(R)\mathscr{S}(R) be the set of all the semisimple complements to JJ in RR. Fix S𝒮(R)S\in\mathscr{S}(R), so that R=SJR=S\oplus J, and write S=i=1NSiS=\bigoplus_{i=1}^{N}S_{i} for some simple rings SiS_{i}, each with unity eie_{i}. For all 1i,jN1\leqslant i,j\leqslant N, define Jij:=eiJejJ_{ij}:=e_{i}Je_{j} to be the (Si,Sj)(S_{i},S_{j})-bimodule obtained via a Peirce decomposition of JJ. Note that J=i,jJijJ=\bigoplus_{i,j}J_{ij}. Let λij\lambda_{ij} be the length of JijJ_{ij} as an (Si,Sj)(S_{i},S_{j})-bimodule; if Jij={0}J_{ij}=\{0\}, then we take λij=0\lambda_{ij}=0.

As in Section 4, our strategy is to show that in a σu\sigma_{u}-elementary ring, most of the bimodules JijJ_{ij} must be zero. Much of the work done earlier in this paper still holds when stated only for unital subrings or σu\sigma_{u}-elementary rings. As noted earlier, Theorem 4.1, and Lemma 4.9 hold mutatis mutandis for unital maximal subrings and σ\sigma-elementary rings (see Remarks 4.6 and 4.10). In particular, we have the following structural results on a σu\sigma_{u}-elementary ring RR.

Lemma 7.5.

Let RR be as in Notation 7.4.

  1. (1)

    If RR is σu\sigma_{u}-elementary, then there is at most one pair (i,j)(i,j) with 1i<jN1\leqslant i<j\leqslant N and Jij{0}J_{ij}\neq\{0\}. For this pair (i,j)(i,j), if Jij{0}J_{ij}\neq\{0\}, then λij=1\lambda_{ij}=1.

  2. (2)

    For all 1kN1\leqslant k\leqslant N, if SkS_{k} is a field and λkk=1\lambda_{kk}=1, then RR is not σu\sigma_{u}-elementary.

Lemma 7.6.

Let RR be a σu\sigma_{u}-elementary ring that is not σ\sigma-elementary.

  1. (1)

    There exists a two-sided ideal II of RR such that JIJ\subseteq I and R/Ii=1p𝔽pR/I\cong\bigoplus_{i=1}^{p}\mathbb{F}_{p}.

  2. (2)

    σ(R)=p+(p2)\sigma(R)=p+\binom{p}{2}.

  3. (3)

    If RR has a residue ring isomorphic to i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}, then R=i=1p+1𝔽pR=\bigoplus_{i=1}^{p+1}\mathbb{F}_{p} and σu(R)=p+(p2)\sigma_{u}(R)=p+\binom{p}{2}.

Proof.

(1) Since RR is not σ\sigma-elementary, it projects nontrivially onto a σ\sigma-elementary ring R/IR/I such that σ(R)=σ(R/I)\sigma(R)=\sigma(R/I). By Proposition 7.3, all σ\sigma-elementary rings are σu\sigma_{u}-elementary except for i=1p𝔽p\bigoplus_{i=1}^{p}\mathbb{F}_{p}. If R/I≇i=1p𝔽pR/I\not\cong\bigoplus_{i=1}^{p}\mathbb{F}_{p}, then σu(R/I)=σ(R/I)=σ(R)σu(R)\sigma_{u}(R/I)=\sigma(R/I)=\sigma(R)\leqslant\sigma_{u}(R), a contradiction to RR being σu\sigma_{u}-elementary. Thus, there exists an ideal II of RR such that R/Ii=1p𝔽pR/I\cong\bigoplus_{i=1}^{p}\mathbb{F}_{p}. Since R/IR/I is semisimple, JIJ\subseteq I.

(2) With II as in part (1), we have σ(R)=σ(R/I)=p+(p2)\sigma(R)=\sigma(R/I)=p+\binom{p}{2}.

(3) Assume that RR has a residue ring isomorphic to i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}. By Lemma 7.2, i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p} is σu\sigma_{u}-elementary and has unital covering number p+(p2)p+\binom{p}{2}. By Lemma 7.1 and part (2), p+(p2)=σ(R)σu(R).p+\binom{p}{2}=\sigma(R)\leqslant\sigma_{u}(R). Since RR is σu\sigma_{u}-elementary, we must have R=i=1p+1𝔽pR=\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}. ∎

Using Lemma 7.6, we can impose another restriction on RR and assume that RR has a residue ring isomorphic to i=1p𝔽p\bigoplus_{i=1}^{p}\mathbb{F}_{p}, but does not have a residue ring isomorphic to i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}. Thus, we may assume that SS has the form

(7.7) S=(i=1pSi)(i=p+1NSi),S=\Big{(}\bigoplus_{i=1}^{p}S_{i}\Big{)}\oplus\Big{(}\bigoplus_{i=p+1}^{N}S_{i}\Big{)},

where Si𝔽pS_{i}\cong\mathbb{F}_{p} for all 1ip1\leqslant i\leqslant p, and Si≇𝔽pS_{i}\not\cong\mathbb{F}_{p} for all p+1iNp+1\leqslant i\leqslant N.

Lemma 7.8.

Let SS be as in (7.7) and let MM be a maximal unital subring of SS.

  1. (1)

    If MM is Type Π1\Pi_{1}, then MM contains i=1pSi\bigoplus_{i=1}^{p}S_{i}.

  2. (2)

    If MM does not contain i=1pSi\bigoplus_{i=1}^{p}S_{i}, then MM is Type Π2\Pi_{2} with SiS_{i} linked to SjS_{j} for some 1i<jp1\leqslant i<j\leqslant p.

Proof.

(1) Assume that MM is Type Π1\Pi_{1}. Then, for some 1iN1\leqslant i\leqslant N,

M=S1Si1MiSi+1SN,M=S_{1}\oplus\cdots\oplus S_{i-1}\oplus M_{i}\oplus S_{i+1}\oplus\cdots\oplus S_{N},

where MiM_{i} is a maximal subring of SiS_{i}. Suppose that 1ip1\leqslant i\leqslant p. Then, Mi={0}M_{i}=\{0\} because Si𝔽pS_{i}\cong\mathbb{F}_{p}. But then, 1RM1_{R}\notin M, a contradiction. So, ip+1i\geqslant p+1, and i=1pSiM\bigoplus_{i=1}^{p}S_{i}\subseteq M.

(2) If i=1pSi\bigoplus_{i=1}^{p}S_{i} is not a subring of MM, then by part (1) MM must be Type Π2\Pi_{2}. In this case, the only way that MM will fail to contain i=1pSi\bigoplus_{i=1}^{p}S_{i} is if SiS_{i} is linked to SjS_{j} for some 1i<jp1\leqslant i<j\leqslant p. ∎

Proposition 7.9.

Let RR be as Notation 7.4, and let SS be as in (7.7). Assume that RR is σu\sigma_{u}-elementary, but not σ\sigma-elementary.

  1. (1)

    λii=0\lambda_{ii}=0 for all 1ip1\leqslant i\leqslant p.

  2. (2)

    Jij={0}J_{ij}=\{0\} for all 1ip1\leqslant i\leqslant p and all p+1jNp+1\leqslant j\leqslant N.

Proof.

(1) Suppose first that λii2\lambda_{ii}\geqslant 2 for some 1ip1\leqslant i\leqslant p. Then, SiJii𝔽p(+)𝔽pλiiS_{i}\oplus J_{ii}\cong\mathbb{F}_{p}(+)\mathbb{F}_{p}^{\lambda_{ii}} (the idealization of 𝔽p\mathbb{F}_{p} with the λii\lambda_{ii}-dimensional vector space 𝔽pλii\mathbb{F}_{p}^{\lambda_{ii}}), which projects onto 𝔽p(+)𝔽p2\mathbb{F}_{p}(+)\mathbb{F}_{p}^{2}. By Lemma 3.2(4), RR has a residue ring isomorphic to 𝔽p(+)𝔽p2\mathbb{F}_{p}(+)\mathbb{F}_{p}^{2}. Using this and Lemmas 7.1 and 7.6, we get

p+(p2)σu(R)σu(𝔽p(+)𝔽p2)=p+1.p+\textstyle\binom{p}{2}\leqslant\sigma_{u}(R)\leqslant\sigma_{u}(\mathbb{F}_{p}(+)\mathbb{F}_{p}^{2})=p+1.

Since RR is σu\sigma_{u}-elementary, this forces R𝔽p(+)𝔽p2R\cong\mathbb{F}_{p}(+)\mathbb{F}_{p}^{2}. But, this contradicts the fact that RR is not σ\sigma-elementary.

So, we must have λii1\lambda_{ii}\leqslant 1 for all 1ip1\leqslant i\leqslant p. However, by Lemma 7.5(2), if λii=1\lambda_{ii}=1 for any 1ip1\leqslant i\leqslant p, then RR is not σu\sigma_{u}-elementary. We conclude that λii=0\lambda_{ii}=0 for all 1ip1\leqslant i\leqslant p.

(2) Suppose by way of contradiction that Jk{0}J_{k\ell}\neq\{0\} for some 1kp1\leqslant k\leqslant p and p+1Np+1\leqslant\ell\leqslant N. By Lemma 7.5(1), there is at most one pair (i,j)(i,j) with 1i<jN1\leqslant i<j\leqslant N and Jij{0}J_{ij}\neq\{0\}. Thus, (k,)(k,\ell) is unique, and JkJ_{k\ell} is a simple (Sk,S)(S_{k},S_{\ell})-bimodule.

Assume without loss of generality that k=pk=p. Let T=i=1p1Sii=1p1𝔽pT=\bigoplus_{i=1}^{p-1}S_{i}\cong\bigoplus_{i=1}^{p-1}\mathbb{F}_{p}. We claim that TT is a two-sided ideal of RR. To see this, note that by part (1), Jii={0}J_{ii}=\{0\} for all 1ip11\leqslant i\leqslant p-1. Hence, J=Jp(i=p+1NJii)J=J_{p\ell}\oplus\big{(}\bigoplus_{i=p+1}^{N}J_{ii}\big{)}. Let T^=(i=pNSi)J\widehat{T}=\big{(}\bigoplus_{i=p}^{N}S_{i}\big{)}\oplus J. Then, R=TT^R=T\oplus\widehat{T}, and elements from TT and T^\widehat{T} mutually annihilate one another. Thus, TT is an ideal of RR. Moreover, the action of 1+J1+J on TT is trivial, so TT is contained in SS^{\prime} for all S𝒮(R)S^{\prime}\in\mathscr{S}(R). We will show that σu(R)=σu(R/T)\sigma_{u}(R)=\sigma_{u}(R/T), which contradicts the fact that RR is σu\sigma_{u}-elementary.

Let 𝒞1\mathcal{C}_{1} be the set of all unital maximal subrings of RR that contain TT, and let 𝒞2\mathcal{C}_{2} be the set of all unital maximal subrings that do not. Let M𝒞2M\in\mathcal{C}_{2}, and suppose that JMJ\not\subseteq M. Then, by [32, Theorem 3.10], M=SIM=S^{\prime}\oplus I for some S𝒮(R)S^{\prime}\in\mathscr{S}(R) and some maximal subideal IJI\subseteq J. But then, TSMT\subseteq S^{\prime}\subseteq M, a contradiction. So, each subring in 𝒞2\mathcal{C}_{2} contains JJ. Use a bar to denote passage to R/JSR/J\cong S. By Lemma 7.8, M¯\overline{M} is Type Π2\Pi_{2} with S¯i\overline{S}_{i} linked to S¯j\overline{S}_{j} for some 1i<jp1\leqslant i<j\leqslant p, and this is true for every M𝒞2M\in\mathcal{C}_{2}.

From here, we will show that for each aT^R/Ta\in\widehat{T}\cong R/T, there exists αT\alpha\in T such that a+αa+\alpha is in a maximal subring containing TT. Let aT^a\in\widehat{T}, and write aa as a=αp+ba=\alpha_{p}+b, where αpSp\alpha_{p}\in S_{p} and b(i=p+1NSi)Jb\in\big{(}\bigoplus_{i=p+1}^{N}S_{i}\big{)}\oplus J. Note that αp\alpha_{p} is equal to an element of 𝔽p\mathbb{F}_{p}. Let α1,,αp1\alpha_{1},\ldots,\alpha_{p-1} be all of the elements of 𝔽p\mathbb{F}_{p} that are not equal to αp\alpha_{p}. Considering αi\alpha_{i} as an element of SiS_{i} for each 1ip11\leqslant i\leqslant p-1, we take α:=i=1p1αiT\alpha:=\sum_{i=1}^{p-1}\alpha_{i}\in T.

Certainly, a+αa+\alpha lies in some unital maximal subring MM of RR. Suppose that M𝒞2M\in\mathcal{C}_{2}. As noted above, M¯\overline{M} is Type Π2\Pi_{2} with S¯i\overline{S}_{i} linked to S¯j\overline{S}_{j} for some 1i<jp1\leqslant i<j\leqslant p. Since the only automorphism of 𝔽p\mathbb{F}_{p} is the identity, this means that αi¯=αj¯\overline{\alpha_{i}}=\overline{\alpha_{j}}. However, this is impossible, because α1,,αp\alpha_{1},\ldots,\alpha_{p} are all distinct elements of 𝔽p\mathbb{F}_{p}. Thus, a+αa+\alpha lies in some subring in 𝒞1\mathcal{C}_{1}.

To complete the proof, let 𝒞\mathcal{C} be a minimal cover of RR by unital subrings, and let 𝒜=𝒞𝒞1\mathcal{A}=\mathcal{C}\cap\mathcal{C}_{1}. Then, each subring in 𝒜\mathcal{A} contains TT, and their images in R/TR/T cover R/TT^R/T\cong\widehat{T}. Hence, σu(R/T)|𝒜|σu(R)\sigma_{u}(R/T)\leqslant|\mathcal{A}|\leqslant\sigma_{u}(R), which contradicts the fact that RR is σu\sigma_{u}-elementary. ∎

Proposition 7.10.

Let RR be as described prior to Lemma 7.6, and let SS be as in (7.7). Assume that RR is σu\sigma_{u}-elementary, but not σ\sigma-elementary. Then,

R=(i=1pSi)(1i,jpJij).R=\Big{(}\bigoplus_{i=1}^{p}S_{i}\Big{)}\oplus\Big{(}\bigoplus_{1\leqslant i,j\leqslant p}J_{ij}\Big{)}.
Proof.

By Proposition 7.9, we know that Jij={0}J_{ij}=\{0\} for all 1ip1\leqslant i\leqslant p and p+1jNp+1\leqslant j\leqslant N. Decompose RR as R=R1R2R=R_{1}\oplus R_{2}, where

R1\displaystyle R_{1} =(i=1pSi)(1i,jpJij), and\displaystyle=\Big{(}\bigoplus_{i=1}^{p}S_{i}\Big{)}\oplus\Big{(}\bigoplus_{1\leqslant i,j\leqslant p}J_{ij}\Big{)},\text{ and }
R2\displaystyle R_{2} =(i=p+1NSi)(p+1i,jNJij).\displaystyle=\Big{(}\bigoplus_{i=p+1}^{N}S_{i}\Big{)}\oplus\Big{(}\bigoplus_{p+1\leqslant i,j\leqslant N}J_{ij}\Big{)}.

We claim that all unital maximal subrings of RR respect this decomposition. That is, if MM is a unital maximal subring of RR, then either M=M1R2M=M_{1}\oplus R_{2} with M1M_{1} maximal in R1R_{1}, or M=R1M2M=R_{1}\oplus M_{2} with M2M_{2} maximal in R2R_{2}. When JMJ\not\subseteq M, this can be shown just as in the proof of Proposition 6.2. When JMJ\subseteq M, M=TJM=T\oplus J for some maximal subring TT of SS. In this case, the only way that MM would not respect the direct sum decomposition of RR is if TT were Type Π2\Pi_{2} with SiS_{i} linked to SjS_{j} for some 1ip1\leqslant i\leqslant p and p+1jNp+1\leqslant j\leqslant N. However, this is impossible, because Sj≇𝔽pS_{j}\not\cong\mathbb{F}_{p} for all p+1jNp+1\leqslant j\leqslant N.

Since every maximal subring of RR has the form M1R2M_{1}\oplus R_{2} or R1M2R_{1}\oplus M_{2}, we can apply Lemma 2.3, which yields σu(R)=min{σu(R1),σu(R2)}\sigma_{u}(R)=\min\{\sigma_{u}(R_{1}),\sigma_{u}(R_{2})\}. Since RR is σu\sigma_{u}-elementary, this means that either R=R1R=R_{1} or R=R2R=R_{2}. Because only R1R_{1} has i=1p𝔽p\bigoplus_{i=1}^{p}\mathbb{F}_{p} as a residue ring, we conclude that R=R1R=R_{1}. ∎

Proposition 7.11.

Let RR be a σu\sigma_{u}-elementary ring that is not σ\sigma-elementary, and such that R≇i=1p+1𝔽pR\not\cong\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}. Then, RA(1,2,2)R\cong A(1,2,2).

Proof.

By Lemma 7.6, RR has no residue ring isomorphic to i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p}. By Proposition 7.10, R=(i=1p𝔽p)JR=\big{(}\bigoplus_{i=1}^{p}\mathbb{F}_{p}\big{)}\oplus J, and J{0}J\neq\{0\} because RR is coverable by unital subrings. By Proposition 7.9, Jii={0}J_{ii}=\{0\} for all 1ip1\leqslant i\leqslant p. Moreover, by Lemma 7.5(1), there is a unique pair (i,j)(i,j) with iji\neq j such that Jij{0}J_{ij}\neq\{0\}, and λij=1\lambda_{ij}=1.

Thus, SiSjJijA(1,p,p)S_{i}\oplus S_{j}\oplus J_{ij}\cong A(1,p,p), which has unital covering number p+1p+1. By 3.2, A(1,p,p)A(1,p,p) occurs as a residue ring of RR. Thus, p+(p2)σu(R)p+1p+\binom{p}{2}\leqslant\sigma_{u}(R)\leqslant p+1, which forces p=2p=2 and σu(R)=3\sigma_{u}(R)=3. Since RR is σu\sigma_{u}-elementary, we must have R=SiSjJijA(1,2,2)R=S_{i}\oplus S_{j}\oplus J_{ij}\cong A(1,2,2). ∎

Proof of Theorem 1.6.

Apply Proposition 7.3, Lemma 7.6, and Proposition 7.11. The covering numbers for i=1p+1𝔽p\bigoplus_{i=1}^{p+1}\mathbb{F}_{p} and A(1,2,2)A(1,2,2) were found in Lemma 7.2. ∎

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

(N):={m:mN,σ(R)=m for some ring R}.\mathscr{E}(N):=\{m:m\leqslant N,\sigma(R)=m\text{ for some ring }R\}.

Let logx\log x denote the binary (base 2) logarithm. Our goal is to show that |(N)||\mathscr{E}(N)| is bounded above by cN/logNcN/\log N for some positive constant cc. Rather than attempting to find an optimal value for cc, we will be content with a value that is easy to verify.

Let π\pi be the prime counting function, which counts the number of primes less than or equal to a given positive number. The following bounds on π(x)\pi(x) will be useful in later estimates.

Lemma 8.1.

For all x>1x>1, π(x)<2x/logx\pi(x)<2x/\log x. For all x>5x>5, x/logx<π(x)x/\log x<\pi(x).

Proof.

Let lnx\ln x denote the natural logarithm. By [29, Corollary 1], we have π(x)<(1.25506x)/(lnx)\pi(x)<(1.25506x)/(\ln x) when x>1x>1, and π(x)>x/ln(x)\pi(x)>x/\ln(x) when x17x\geqslant 17. The remaining cases for the lower bound follow from inspection. ∎

Let Π\Pi be the prime power counting function, i.e.,

Π(x):=|{m:2mx and m is a prime power}|.\Pi(x):=|\{m\in\mathbb{N}:2\leqslant m\leqslant x\text{ and }m\text{ is a prime power}\}|.

Using Lemma 8.1, it is not difficult to get an upper bound on Π(x)\Pi(x).

Lemma 8.2.

Let x>1x>1. Then, Π(x)<8x/logx\Pi(x)<8x/\log x.

Proof.

Let :=logx\ell:=\lfloor\log x\rfloor. Then,

Π(x)\displaystyle\Pi(x) =r=1π(x1/r)=π(x)+r=2π(x1/r)π(x)+(1)π(x1/2)\displaystyle=\sum_{r=1}^{\ell}\pi(x^{1/r})=\pi(x)+\sum_{r=2}^{\ell}\pi(x^{1/r})\leqslant\pi(x)+(\ell-1)\pi(x^{1/2})
2xlogx+(1)2x1/2(1/2)log(x)2xlogx+4x1/2<8xlogx.\displaystyle\leqslant\dfrac{2x}{\log x}+(\ell-1)\dfrac{2x^{1/2}}{(1/2)\log(x)}\leqslant\dfrac{2x}{\log x}+4x^{1/2}<\dfrac{8x}{\log x}.

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 N>1N>1 be a natural number.

  1. (1)

    The number of integers mm, 2mN2\leqslant m\leqslant N, that can be expressed as τ(q)ν(q)+d(τ(q)2)\tau(q)\nu(q)+d\binom{\tau(q)}{2} for some prime power q=pdq=p^{d} is less than 8N/logN8N/\log N.

  2. (2)

    The number of integers mm, 2mN2\leqslant m\leqslant N, that can be expressed as q+1q+1 for some prime power q=pdq=p^{d} is less than 8N/logN8N/\log N.

  3. (3)

    The number of integers mm, 2mN2\leqslant m\leqslant N, that can be expressed as

    1ak=1,akn1(qnqk)+k=1,akn/2(nk)q,\frac{1}{a}\prod_{k=1,\\ a\nmid k}^{n-1}(q^{n}-q^{k})+\sum_{k=1,\\ a\nmid k}^{\lfloor n/2\rfloor}\binom{n}{k}_{q},

    where qq is some prime power, n2n\geqslant 2, aa is the smallest prime divisor of nn, and (nk)q\binom{n}{k}_{q} is the qq-binomial coefficient is less than 56N/logN56N/\log N.

  4. (4)

    Let N>1N>1 be a natural number. The number of integers mm, 2mN2\leqslant m\leqslant N, that can be expressed as

    qn+(nd)q1+ω(d),q^{n}+\binom{n}{d}_{q_{1}}+\omega(d),

    where qq and q1q_{1} are prime powers, q=q1dq=q_{1}^{d}, n3n\geqslant 3, and d<nd<n is less than 72N/logN72N/\log N.

Proof.

The first two parts follow from the fact that there is one such integer mm of the desired form for each prime power. Hence, the number of such integers mm is bounded above by Π(N)\Pi(N), which is less than 8N/logN8N/\log N by Lemma 8.2.

For (3), let

P:=1ak=1,akn1(qnqk) and S:=k=1,akn/2(nk)q.P:=\frac{1}{a}\prod_{k=1,\\ a\nmid k}^{n-1}(q^{n}-q^{k})\;\text{ and }\;S:=\sum_{k=1,\\ a\nmid k}^{\lfloor n/2\rfloor}\binom{n}{k}_{q}.

We first note that NP+S12q2N\geqslant P+S\geqslant\tfrac{1}{2}q^{2}, from which we conclude that q2Nq\leqslant\sqrt{2N}, and there at most Π(2N)\Pi(\sqrt{2N}) choices for qq.

Next, by [33, Lemma 3.31] we have Pqn(n(n/a)1)P\geqslant q^{n(n-(n/a)-1)}, so

NP+S>qn(n(n/a)1)qn(n2)/2q(n2)2/2.N\geqslant P+S>q^{n(n-(n/a)-1)}\geqslant q^{n(n-2)/2}\geqslant q^{(n-2)^{2}/2}.

It follows that (n2)2<2logN(n-2)^{2}<2\log N and n<2+2logNn<2+\sqrt{2\log N}. Since there is exactly one integer expressible as P+SP+S for each pair (q,n)(q,n), the number of such integers at most NN having this form is bounded above by

Π(2N)(2+2logN)<82N(2+2logN)(1/2)log(2N)<56NlogN.\Pi(\sqrt{2N})\cdot(2+\sqrt{2\log N})<\frac{8\sqrt{2N}(2+\sqrt{2\log N})}{(1/2)\log(2N)}<\frac{56N}{\log N}.

Finally, for (4), since q3qn<Nq^{3}\leqslant q^{n}<N, there are at most Π(N3)\Pi(\sqrt[3]{N}) choices for qq, and, since 2nqn<N2^{n}\leqslant q^{n}<N and d<nd<n, there are at most logN\log N choices for nn for a fixed qq, and at most logN\log N choices for dd given a fixed nn. Hence, the number of integers expressible in the desired form is bounded above by the number of triples (q,n,d)(q,n,d), which is bounded above by

Π(N3)(logN)2<8N3(logN)2(1/3)logN<72NlogN.\Pi(\sqrt[3]{N})(\log N)^{2}<\frac{8\sqrt[3]{N}(\log N)^{2}}{(1/3)\log N}<\frac{72N}{\log N}.

We now have what we need to prove Corollary 1.9.

Proof of Corollary 1.9.

The upper bound of 144N/logN144N/\log N follows from Lemma 8.3. For the lower bound, we note that by Theorem 1.3 (2), every integer of the form q+1q+1, where qq is a prime power, is a covering number of a ring with unity. Thus, |(N)|Π(N)1|\mathscr{E}(N)|\geqslant\Pi(N)-1, and for N5N\geqslant 5 we have Π(N)1π(N)>N/logN\Pi(N)-1\geqslant\pi(N)>N/\log N. ∎

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 nn-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 nn-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 σ\sigma-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.