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

Adjacency-Like Conditions and Induced Ideal Graphs

Saba Al-Kaseasbeh Department of Mathematics
Tafila Technical University
Tafilah, Jordan
[email protected]
 and  Jim Coykendall Department of Mathematical Sciences
Clemson University
Clemson, SC
[email protected]
Abstract.

In this paper we examine some natural ideal conditions and show how graphs can be defined that give a visualization of these conditions. We examine the interplay between the multiplicative ideal theory and the graph theoretic structure of the associated graph. Behavior of these graphs under standard ring extensions are studied, and in conjunction with the theory, some classical results and connections are made.

Key words and phrases:
Noetherian, SFT
2010 Mathematics Subject Classification:
Primary: 13C05, 13A15, 13C13

1. Introduction and Background

Recently there has been much attention paid to various aspects of commutative algebra that lend themselves to a graph-theoretic approach. The origins of some of the aspects of this recent interest in the interplay between graphs and commutative algebra can perhaps be traced (chronologically) to the papers [6] and [2]. The Anderson-Livingston paper defined the notion of a zero divisor graph that is more commonly used today, and it has motivated an industry of work on the interplay between commutative algebra and graph theory. We recall that a zero divisor graph of a commutative ring, RR, has vertex set defined to be the set of nonzero zero divisors, and it is declared that there is an edge between a,bRa,b\in R if and only if ab=0ab=0. In [2] the following remarkable theorem is proved:

Theorem 1.1.

Let RR be a commutative ring with identity and Z(R)Z(R) its zero divisor graph. Then Z(R)Z(R) is connected and has diameter less than or equal to 3.

The authors consider this theorem far from obvious at first blush and find it amazing that any commutative ring with identity must have such a well-behaved structure for its zero divisor graph.

These works have inspired myraid questions and lines of study from numerous authors. For example, the concept of an irreducible divisor graph was introduced by the second author and J. Maney in [9]. Among other things, it is shown that many factorization properties, like unique factorization or the half-factorial property, can be seen in the properties of the irreducible divisor graph(s). Additionally, variants of the zero divisor graph have been studied by numerous authors, for example [9] and [8], and [1]. Of these three, [8] has particular relevance to the theme of this paper, as in this work, subtle variants of atomicity in integral domains are revealed by the connectedness of an associated graph. Other studies have focused on various properties of zero-divisor graphs and their variants (for example, the interested reader can consult [4], [5], [17] among others).

Aside from zero divisor graphs, there has been much other work done in the intersection of commutative algebra and graph theory since the 1990’s. In the influential paper [20], Cohen-Macaulay graphs were studied and (among other things) a strong connection was made between the unmixedness of an ideal of a polynomial ring associated to a complex and the Cohen-Macaulay property of the ring modulo this ideal. In [18] Rees algebras are studied with the aid of (and in parallel to) combinatorial properties of graphs naturally associated to the rings in question, and in [21] edge ideals were developed and explored. A couple of more recent works (among many) that deal with edge ideals are [13] and [14]. In short, combinatorial methods in commutative algebra, and in particular the interplay between graph theory and commutative algebra, has borne much fruit in its relatively short history and continues to be a very fertile area of research.

In this work, we adopt a slightly different approach; the motivation is to gain insight into the ideal structure of a commutative ring by assigning a graph to the set of ideals (sometimes with natural restrictions or slight variants). Three types of graphs are developed: the first types measure adjacency or “closeness” of the ideals in RR, the second types mimic the structure of the ideal class group and the zero divisor graphs of Anderson/Livingston, and the third types measure finite containment of ideals.

In the third section, the adjacency graphs are given and strong connections are given with respect to the connectedness of the graphs. In particular, a complete classification is given for connectedness for all graphs and strong bounds are given for the diameters.

In the fourth section, graphs are defined that emulate the behavior of the class group of RR, where RR is an integral domain. The connected/complete case is resolved and in the case that RR is a Dedekind domain, the connected components of the graph are studied.

In the fifth section, a graph generalizing the classical zero-divisor graph is introduced and (among other things), we capture Theorem 1.1 as a corollary.

In the sixth and final section, we explore the notion of “finite containment” graphs. As it turns out, these graphs are almost always connected (the exception being the case where RR is quasilocal with maximal ideal 𝔐\mathfrak{M} that is not finitely generated). Here the study of the diameter turns out to be a fruitful line of attack and the behavior of diameter with regard to standard extensions (polynomial and power series) as well as homomorphic images are explored. Throughout all sections, a number of examples designed to illuminate are presented.

As a final note, we introduce some notation to be used. If II and JJ are vertices of a graph, we will use the notation IJI-J to indicate an edge between II and JJ. The notation IJI\leftrightarrow J will signify that either I=JI=J or that there is an edge between II and JJ. This notation will prove convenient in the sequel.

2. Ideal Theoretic Graphs

In this section we define and justify some types of graphs that are determined by ideal theoretic properties. We wish for our graphs to be simple (no loops or multiple edges), and so in the sequel when we say that there is an edge between the vertices uu and vv, we default to the condition that uu and vv are distinct.

In general we consider, XX, a collection of ideals of a commutative ring, RR, as our collection of vertices (in many cases XX will consist of all ideals or all nonzero ideals of RR). To determine an edge set we let P(,)P(-,-) be a statement and we say that there is an edge between distinct II and JJ if and only if P(I,J)P(I,J) is true.

We now present three definitions that explore variants of the types of ideal graphs that we investigate. The first set of definitions measure “closeness” of the ideals of RR in a certain sense; they are designed to measure/highlight the density of the ideal structure of RR.

Definition 2.1 (Adjacency Graphs).

Let RR be a commutative ring with identity. We define the following graphs associated to RR. In all cases, the vertex set is the set of proper ideals of RR, and if the ideals II and JJ have an edge between them, then IJI\neq J.

  1. (0)

    In GA0(R)GA_{0}(R), we say that II and JJ have an edge between them if and only if II and JJ are adjacent ideals.

  2. (1)

    In GA1(R)GA_{1}(R), we say that II and JJ have an edge between them if and only if there is a maximal ideal 𝔐\mathfrak{M} such that I=J𝔐I=J\mathfrak{M}.

  3. (2)

    In GA2(R)GA_{2}(R), we say that II and JJ have an edge between them if and only if (I:RJ)=𝔐(I:_{R}J)=\mathfrak{M} is a maximal ideal.

The next definitions are designed to reflect the notions of ideal multiplication and class structure. The graphs GClint(R)GCl_{int}(R) and GCl(R)GCl(R) are designed to graphically represent variants of the ideal class group. GZ(R)GZ(R) is an ideal-theoretic analogue of the zero-divisor graph of Anderson-Livingston. GZ(R)GZ(R) with the 00-ideal removed has been considered in [7] where the terminology “annihilating-ideal graph” is used. We will respect this terminology, but will use the notation provided for ease and uniformity of exposition.

Definition 2.2 (Ideal Multiplication/Class Structure Graphs).

Let RR be a commutative ring with identity. We define the following graphs associated to RR. For these definitions, the vertex set is the set of nonzero ideals of RR and in the case of GZ(R)GZ(R) we also demand that the ideals are proper. As above, if the ideals II and JJ have an edge between them, then IJI\neq J.

  1. (1)

    Let RR be an integral domain. For GClint(R)GCl_{int}(R), we say that II and JJ have an edge between them if and only if there is a nonzero aRa\in R such that I=JaI=Ja.

  2. (2)

    Let RR be an integral domain with quotient field KK. In GCl(R)GCl(R), we say that II and JJ have an edge between them if and only if there is a nonzero kKk\in K such that I=JkI=Jk.

  3. (3)

    In GZ(R)GZ(R) we say that II and JJ have an edge between them if and only if IJ=0IJ=0.

The last set of definitions is concerned with ideals that are finitely generated (resp. principally generated) over some initial ideal. The motivation here is to investigate possible paths between ideals in a commutative ring where the steps can be done “one generator (or a finite number of generators) at a time.” Paths between two vertices (ideals) II and JJ of these graphs rely heavily on the interplay between the maximal ideals containing II and the maximal ideals containing JJ. As a consequence, we will see that these graphs give insight to the structure of MaxSpec(R)={𝔐R|𝔐 is maximal}\text{MaxSpec}(R)=\{\mathfrak{M}\subset R|\mathfrak{M}\text{ is maximal}\}.

Definition 2.3 (Finite Containment Graphs).

Let RR be a commutative ring with identity. We define the following graphs associated to RR; in all cases, the vertex set is the set of proper ideals of RR.

  1. (1)

    In GF(R)GF(R) we say that II and JJ have an edge between them if and only if IJI\subset J and J/IJ/I is a finitely generated ideal of R/IR/I.

  2. (2)

    In GP(R)GP(R) we say that II and JJ have an edge between them if and only if IJI\subset J and J/IJ/I is a principal ideal of R/IR/I.

Remark 2.4.

For all of the graphs defined above (with the exceptions of GA2(R)GA_{2}(R), GCl(R)GCl(R), and GZ(R)GZ(R)) containment is used or implied in the definition. Hence one could attempt to glean more information by directing the graph, but we will take no notice of this possibility in the present work.

Remark 2.5.

We note that if, in the definition of GClint(R)GCl_{int}(R), aa is assumed to be irreducible, then this has been investigated by J. Boynton and the second author [8].

3. Adjacency: The Graphs GA0(R),GA1R,and GA2(R)GA_{0}(R),GA_{1}{R},\text{and }GA_{2}(R)

In this section, we look at the graphs that mirror adjacency; as we will see the concepts of adjacency, maximal conductor, and maximal multiple are intimately related. The metrics here, in a certain sense, strive to measure how “tightly packed” the ideals of RR are. Additionally, there is a certain measure of discreteness being measured when considering adjacency conditions.

We begin by developing preliminary lemmata that will be essential later. We first note the relationship between the defining concepts of GA1(R)GA_{1}(R) and GA2(R)GA_{2}(R).

Proposition 3.1.

Let II and JJ be ideals of RR with JIJ\not\subseteq I and 𝔐\mathfrak{M} a maximal ideal. The following are equivalent.

  1. (1)

    (I:RJ)=𝔐(I:_{R}J)=\mathfrak{M}.

  2. (2)

    J𝔐IJ\mathfrak{M}\subseteq I.

Proof.

Assuming condition (1), if x𝔐x\in\mathfrak{M} then xJI\ xJ\subseteq I and hence 𝔐JI\mathfrak{M}J\subseteq I.

On the other hand, if J𝔐IJ\mathfrak{M}\subseteq I then for all x𝔐,JxIx\in\mathfrak{M},\ Jx\subseteq I and hence x(I:RJ)x\in(I:_{R}J). Since we now have that 𝔐(I:RJ)\mathfrak{M}\subseteq(I:_{R}J) and that 𝔐\mathfrak{M} is maximal and JIJ\not\subseteq I, then 𝔐=(I:RJ)\mathfrak{M}=(I:_{R}J). ∎

Lemma 3.2.

Let I,JRI,J\subseteq R be ideals with JIJ\not\subseteq I. If I=J𝔐I=J\mathfrak{M} where 𝔐\mathfrak{M} is a maximal ideal then (I:RJ)=𝔐(I:_{R}J)=\mathfrak{M}, but not necessarily conversely. Hence GA1(R)GA_{1}(R) is a subgraph of GA2(R)GA_{2}(R).

Proof.

Note that 𝔐JI\mathfrak{M}J\subseteq I, and hence 𝔐(I:RJ)\mathfrak{M}\subseteq(I:_{R}J). Since JIJ\not\subseteq I, we must have equality.

To see that the converse does not hold in general, consider the domain K[x,y]K[x,y], where KK is any field, and the ideals I=(x,xy,y2)I=(x,xy,y^{2}) and J=(x,y)J=(x,y). Note that (I:RJ)=J(I:_{R}J)=J but J2IJ^{2}\subsetneq I. ∎

Lemma 3.3.

If IJI\subsetneq J are adjacent, then (I:RJ)=𝔐(I:_{R}J)=\mathfrak{M} is maximal, but not necessarily conversely. Hence GA0(R)GA_{0}(R) is a subgraph of GA2(R)GA_{2}(R).

Proof.

Let IJI\subsetneq J be adjacent and let 𝒞=(I:RJ)\mathcal{C}=(I:_{R}J); we will show that 𝒞\mathcal{C} is maximal. Since II is strictly contained in JJ, we can find an xJIx\in J\setminus I; additionally, we note that J=(I,x)J=(I,x) because of the adjacency of II and JJ.

We now choose an arbitrary z𝒞z\notin\mathcal{C} and note that zxIzx\notin I (indeed, if zxIzx\in I then the fact that J=(I,x)J=(I,x) would show that z𝒞z\in\mathcal{C} which is a contradition). Hence, it must be the case that J=(I,zx)J=(I,zx), and since xJx\in J, we obtain

x=rzx+αx=rzx+\alpha

for some rRr\in R and αI\alpha\in I. Rearranging the above, we now have

x(1rz)=αI.x(1-rz)=\alpha\in I.

Since (1rz)(1-rz) conducts xx to II and J=(I,x)J=(I,x), (1rz)𝒞(1-rz)\in\mathcal{C}. Therefore (𝒞,z)=R(\mathcal{C},z)=R for all z𝒞z\notin\mathcal{C} and so 𝒞\mathcal{C} is maximal.

To see that the converse does not hold, we revisit the example of the previous result. Letting R=K[x,y]R=K[x,y], A=(x2,xy,y2),I=(x,xy,y2),J=(x,y)A=(x^{2},xy,y^{2}),I=(x,xy,y^{2}),J=(x,y). Note that (A:RJ)=J(A:_{R}J)=J but AA and JJ are not adjacent as AIJA\subsetneq I\subsetneq J. ∎

Lemma 3.4.

Suppose that IJI\subsetneq J are adjacent and 𝒞:=(I:RJ)\mathcal{C}:=(I:_{R}J). If 𝒞J=I\mathcal{C}\bigcap J=I then JIJ\setminus I is a multiplicatively closed set.

Proof.

Let x,yJIx,y\in J\setminus I. Certainly xyJxy\in J. By way of contradiction, we assume that xyIxy\in I. Since II and JJ are adjacent and xIx\notin I, (x,I)=J(x,I)=J.

Now let jJj\in J be arbitrary. By the previous remark, we can find rRr\in R and iIi\in I such that j=rx+ij=rx+i. Multiplying this by yy we obtain that yj=rxy+iyIyj=rxy+iy\in I. Hence y𝒞J=Iy\in\mathcal{C}\bigcap J=I which is the desired contradiction. ∎

Lemma 3.5.

Let IJI\subsetneq J be adjacent and ARA\subseteq R an ideal. Then the ideals IAI\bigcap A and JAJ\bigcap A are either equal or adjacent.

Proof.

We will assume that IAI\bigcap A and JAJ\bigcap A are distinct and suppose that x(JA)(IA)x\in(J\bigcap A)\setminus(I\bigcap A). Since xIx\notin I, it must be the case that J=(I,x)J=(I,x).

Let jJAj\in J\bigcap A be arbitrary. Since J=(I,x)J=(I,x), we have that

j=rx+αj=rx+\alpha

for some rRr\in R and αI\alpha\in I. We observe further that rxJArx\in J\bigcap A, and hence, αIA\alpha\in I\bigcap A. We conclude that

JA=(IA,x)J\bigcap A=(I\bigcap A,x)

for any x(JA)(IA)x\in(J\bigcap A)\setminus(I\bigcap A), and so IAI\bigcap A and JAJ\bigcap A must be adjacent. ∎

Lemma 3.6.

If IJI\subsetneq J are adjacent ideals and xRx\in R, then the ideals (I,x)(I,x) and (J,x)(J,x) are either equal or adjacent.

Proof.

Suppose that AA is an ideal with (I,x)A(J,x)(I,x)\subsetneq A\subseteq(J,x) and let aA(I,x)a\in A\setminus(I,x). We write a=j+rxa=j+rx with jJIj\in J\setminus I. Since II and JJ are adjacent and jIj\notin I, (I,j)=J(I,j)=J. As AA contains xx, AA contains all of JJ; we conclude that A=(J,x)A=(J,x) and the proof is complete. ∎

Proposition 3.7.

Let IJI\subsetneq J be adjacent ideals and SRS\subset R be a multiplicatively closed set (not containing 0). Then the ideals ISJSI_{S}\subseteq J_{S} are either adjacent or equal.

Proof.

Since IJI\subsetneq J are adjacent, there is an xJIx\in J\setminus I such that J=(I,x)J=(I,x) (in fact, any xJIx\in J\setminus I will do). Suppose that ISJSI_{S}\subsetneq J_{S} and let bsJSIS\frac{b}{s}\in J_{S}\setminus I_{S}. We can assume that bJIb\in J\setminus I and since IJI\subsetneq J are adjacent, we can find an rRr\in R and αI\alpha\in I such that b=rx+αb=rx+\alpha. In RSR_{S} we have the equation

bs=(rs)(x1)+αs\frac{b}{s}=(\frac{r}{s})(\frac{x}{1})+\frac{\alpha}{s}

which shows that JSJ_{S} is generated over ISI_{S} by any element of the form x1\frac{x}{1} with xJIx\in J\setminus I. Hence ISI_{S} and JSJ_{S} are either equal or adjacent. ∎

The next result is a consequence of the Lattice Isomorphism Theorem and the one after is straightforward; we omit the proofs.

Proposition 3.8.

Let A,BRA,B\subseteq R be ideals containing the ideal II. Then AA and BB are adjacent in RR if and only if A/IA/I and B/IB/I are adjacent in R/IR/I.

Lemma 3.9.

Let RR be a ring and IJI\subseteq J be ideals. If II is finitely generated and J/IR/IJ/I\subseteq R/I is finitely generated, then JJ is finitely generated.

Proposition 3.10.

Let RR be a 11-dimensional domain. RR is Noetherian if and only if R/IR/I is Artinian for each ideal 0IR0\neq I\subseteq R.

Proof.

Suppose first that RR is Noetherian. If IRI\subseteq R is a nonzero ideal, then R/IR/I is Noetherian of dimension 0 and hence is Artinian.

Now suppose that R/IR/I is Artinian for each nonzero ideal IRI\subset R. It suffices to show that every ideal of RR is finitely generated. Let JRJ\subset R be an arbitrary nonzero ideal. Let 0xJ0\neq x\in J and note that by hypothesis, R/(x)R/(x) is Artinian and so J/(x)J/(x) is finitely generated. From Lemma 3.9 it follows that JJ is finitely generated and hence RR is Noetherian. ∎

We now apply these preliminary results to the ideal graphs in question. We begin with the following result that records the stability of Gi(R)G_{i}(R) for 0i20\leqslant i\leqslant 2 with respect to localization and homomorphic images.

Theorem 3.11.

Let IRI\subset R be an ideal and SRS\subset R be a multiplicatively closed set. If GAi(R)GA_{i}(R) is connected or complete for 0i20\leqslant i\leqslant 2, then so are GAi(RS)GA_{i}(R_{S}) and GAi(R/I)GA_{i}(R/I).

Proof.

The case for GA0()GA_{0}(-) follows immediately from Proposition 3.7 and Proposition 3.8.

For GA1()GA_{1}(-) we suppose that A,BRA,B\subset R are distinct and have an edge between them. Then A=B𝔐A=B\mathfrak{M}, with 𝔐\mathfrak{M} maximal. Note that if 𝔐S\mathfrak{M}\bigcap S\neq\emptyset then 𝔐S=RS\mathfrak{M}_{S}=R_{S} and if not then 𝔐SRS\mathfrak{M}_{S}\subset R_{S} is maximal. As AS=BS𝔐SA_{S}=B_{S}\mathfrak{M}_{S}, we have that ASA_{S} and BSB_{S} are either equal or have an edge between them. Since all ideals of RSR_{S} are of this form, if GA1(R)GA_{1}(R) is either connected or complete, then so if GA1(RS)GA_{1}(R_{S}). Additionally, if IA,BI\subseteq A,B then A/I=(B/I)(𝔐/I)A/I=(B/I)(\mathfrak{M}/I) and from this the result follows for GA1(R/I)GA_{1}(R/I).

In a similar fashion, if we consider GA2()GA_{2}(-) with the ideals/notation in the previous paragraph, we assume that (A:RB)=𝔐(A:_{R}B)=\mathfrak{M}. In this case, it is easy to verify that (A/I:RB/I)=𝔐/I(A/I:_{R}B/I)=\mathfrak{M}/I and 𝔐S(AS:RBS)\mathfrak{M}_{S}\subseteq(A_{S}:_{R}B_{S}). Hence, again, the result follows for GA2()GA_{2}(-). ∎

We remark that the previous result will be an immediate consequence of later structure theorems concerning GAi(R)GA_{i}(R) (see Theorem 3.13 and Theorem 3.16).

The following result is of some independent interest and will be crucial in later work where we show the connection between GA0(R)GA_{0}(R) and the Artinian condition. The result shows that if IJI\subseteq J are ideals and if there is a finite chain of adjacent ideals connecting II and JJ, this chain can be refined to a finite increasing chain.

Proposition 3.12.

Let IJI\subseteq J be ideals and {In}n=0N\{I_{n}\}_{n=0}^{N} ideals with I=I0I=I_{0}, J=INJ=I_{N}, and IkI_{k} and Ik+1I_{k+1} adjacent for each 0kN10\leqslant k\leqslant N-1. Then there exists an increasing chain of ideals

I=J0J1JM=JI=J_{0}\subseteq J_{1}\subseteq\cdots\subseteq J_{M}=J

with each successive pair of ideals adjacent and MNM\leqslant N.

Proof.

Using the notation above, we say that the ideal IkI_{k} is a hinge ideal if IkI_{k} either properly contains both Ik1I_{k-1} and Ik+1I_{k+1} or is properly contained in them both. We proceed by induction on mm, the number of hinge ideals between II and JJ.

At the outset, we simplify matters by assuming that all of the ideals {Is}\{I_{s}\} are contained in JJ by intersecting the collection with JJ and applying Lemma 3.5; we also note that if this chain is increasing, then the conclusion holds. We now describe a reduction process that will greatly simplify the inductive argument.

We first suppose that the first hinge ideal is contained in II; that is, we have the decreasing chain of ideals

I:=I0I1It.I:=I_{0}\supseteq I_{1}\supseteq\cdots\supseteq I_{t}.

Because of the adjacency of successive elements of the chain, we can find xsIsIs+1, 0st1x_{s}\in I_{s}\setminus I_{s+1},\ 0\leqslant s\leqslant t-1 such that

Is=(Is+1,xs).I_{s}=(I_{s+1},x_{s}).

We now make a preliminary refinement of the collection {Is}s=0N\{I_{s}\}_{s=0}^{N} by letting Is=(Is,x0,x1,,xt1)I_{s}^{\prime}=(I_{s},x_{0},x_{1},\cdots,x_{t-1}) for 0sN0\leqslant s\leqslant N. Applying Lemma 3.6 tt times to the entire collection of ideals shows that the new collection {Is}\{I_{s}^{\prime}\} begins as

I=I0=I1==ItIt+1I=I_{0}^{\prime}=I_{1}^{\prime}=\cdots=I_{t}^{\prime}\subseteq I_{t+1}^{\prime}

and the ideals {Is}s=tN\{I_{s}^{\prime}\}_{s=t}^{N} are successively adjacent (or equal). Additionally if IjI_{j} is an ideal in the collection that is not initially a hinge ideal, then IjI_{j}^{\prime} can neither be properly contained in, nor properly contain, both the ideals Ij1I_{j-1}^{\prime} and Ij+1I_{j+1}^{\prime}. So when we identify ideals that are equal, our refined collection has less than or equal to the initial number of ideals and at least one less hinge ideal.

With this in hand, we proceed to argue inductively on the number, mm, of hinge ideals appearing between II and JJ. If m=0m=0 then the chain is increasing and the conclusion is immediate, and this, coupled with the above argument, gives the case m=1m=1. We now suppose that the conclusion holds for m0m\geqslant 0 and consider the case of m+1m+1 hinge ideals.

In the first case, we assume that the first hinge ideal ItI_{t} is contained in II. The previous argument shows that we can refine so that I=ItI=I_{t}^{\prime}, and as above, the number of ideals in the new collection (after equal ideals are identified) is nonincreasing and the number of hinge ideals in the new collection is less than or equal to mm. By induction, we are done in this case.

In the second case, we suppose that the first hinge ideal ItI_{t} contains II. In this case, we consider the collection {Is}s=tN\{I_{s}\}_{s=t}^{N}. Since this collection has mm hinge ideals with the first being contained in ItI_{t}, we apply the first case and extract the increasing chain {Is}s=0k\{I_{s}^{\prime}\}_{s=0}^{k} with I0=ItI_{0}^{\prime}=I_{t} with kNtk\leqslant N-t. Combining this increasing chain with {Ir}r=0t\{I_{r}\}_{r=0}^{t} (and subtracting 11 for the repetition of ItI_{t}), we obtain our increasing chain with length t+k+1Nt+1+t=N+1t+k+1\leqslant N-t+1+t=N+1. ∎

We now characterize some types of rings via their induced adjacency graphs; we begin with GA0()GA_{0}(-).

Theorem 3.13.

Let RR be a commutative ring with identity and GA0(R)GA_{0}(R) its adjacency graph. The following are equivalent.

  1. (1)

    RR is Artinian.

  2. (2)

    GA0(R)GA_{0}(R) is connected.

Proof.

If RR is Artinian, then every finitely generated RR-module (in particular, an ideal of RR) has a composition series. Since every ideal of RR is connected to (0)(0), we have that GA0(R)GA_{0}(R) is connected.

On the other hand, if GA0(R)GA_{0}(R) is connected, then there is a (finite) path between (0)(0) and any ideal of RR. Proposition 3.12 allows this to be refined to a composition series, and hence RR is Artinian. ∎

We now produce the following corollary with a slight variant. We define the graph GA0(R)GA^{*}_{0}(R) to be the subgraph of GA0(R)GA_{0}(R) with the 00-ideal vertex removed.

Corollary 3.14.

Let RR be an integral domain that is not a field. GA0(R)GA^{*}_{0}(R) is connected if and only if RR is Noetherian and dim(R)=1\text{dim}(R)=1.

Proof.

Suppose that RR is 11-dimensional and Noetherian and let I,JRI,J\subseteq R be nonzero ideals. To show that GA0(R)GA^{*}_{0}(R) is connected, it suffices to show that II and IJI\bigcap J can be connected in a finite sequence of steps. To this end, we note that since IJI\bigcap J is nonzero, the ring R/(IJ)R/(I\bigcap J) is 00-dimensional and Noetherian, and hence Artinian. By Theorem 3.13 there is a finite sequence of adjacent ideals (of the displayed form)

(IJ)/(IJ)I1/(IJ)I2/(IJ)I/(IJ)(I\bigcap J)/(I\bigcap J)\subset I_{1}/(I\bigcap J)\subset I_{2}/(I\bigcap J)\subset\cdots\subset I/(I\bigcap J)

connecting I/(IJ)I/(I\bigcap J) to the zero ideal in R/(IJ)R/(I\bigcap J). By Proposition 3.8 this corresponds to a chain of adjacent ideals

(IJ)I1I2I(I\bigcap J)\subset I_{1}\subset I_{2}\subset\cdots\subset I

in RR and hence GA0(R)GA^{*}_{0}(R) is connected.

Now we suppose that GA0(R)GA^{*}_{0}(R) is connected. Let IRI\subset R be an arbitrary nonzero ideal. Proposition 3.8 assures us that adjacency is preserved modulo II and so we obtain that GA0(R/I)GA_{0}(R/I) is connected. Hence Theorem 3.13 gives us that R/IR/I is Artinian (for any nonzero ideal II). From Proposition 3.10 we obtain that RR is Noetherian.

To see that RR is 11-dimensional, we suppose that there is a chain of primes (0)𝔓𝔐(0)\subsetneq\mathfrak{P}\subsetneq\mathfrak{M}. If we choose the ideal I=𝔓I=\mathfrak{P}, we would have that dim(R/𝔓)1\text{dim}(R/\mathfrak{P})\geqslant 1 and hence R/𝔓R/\mathfrak{P} is not Artinan. We conclude that GA0(R/𝔓)GA_{0}(R/\mathfrak{P}) is not connected and again Proposition 3.8 gives us that GA0(R)GA^{*}_{0}(R) cannot be connected and so dim(R)1\text{dim}(R)\leqslant 1. Because of the fact that RR is a domain that is not a field, then the dimension of RR is precisely 11. ∎

Proposition 3.15.

If GA0(R)GA_{0}(R) or GA1(R)GA_{1}(R) is connected, then so is GA2(R)GA_{2}(R).

Proof.

Since the vertex sets are the same, this is immediate from Lemma 3.2 and Lemma 3.3. ∎

We now present a strong characterization of connectedness for the graphs GAi(R)GA_{i}(R), i=1,2i=1,2. Of course this condition is weaker than the Artinian condition.

Theorem 3.16.

Let RR be a commutative ring with identity. The following conditions are equivalent.

  1. (1)

    GA1(R)GA_{1}(R) is connected.

  2. (2)

    GA2(R)GA_{2}(R) is connected.

  3. (3)

    There is a collection of not necessarily distinct maximal ideals {𝔐1,𝔐2,,𝔐n}\{\mathfrak{M}_{1},\mathfrak{M}_{2},\cdots,\mathfrak{M}_{n}\} such that 𝔐1𝔐2𝔐n=0\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0.

Proof.

By Proposition 3.15, we have the implication (1)(2)(1)\Longrightarrow(2). For the implication (2)(3)(2)\Longrightarrow(3), we suppose that GA2(R)GA_{2}(R) is connected and that 𝔐\mathfrak{M} is a maximal ideal of RR. By assumption, there is a finite path from 𝔐\mathfrak{M} to the ideal (0)(0):

𝔐=I0I1I2ImIm+1=0\mathfrak{M}=I_{0}\supsetneq I_{1}\bowtie I_{2}\bowtie\cdots\bowtie I_{m}\supsetneq I_{m+1}=0

where each \bowtie denotes either \supsetneq or \subsetneq. In the proof of this implication, we will use the notion of hinge ideals introduced in the proof of Proposition 3.12. Note that there must be an even number of hinge ideals in the path described above, which we will denote H1,H2,,H2tH_{1},H_{2},\cdots,H_{2t}. So an abbreviated version of the path described above can be expressed in the form

𝔐H1H2H2t0\mathfrak{M}\supsetneq H_{1}\subsetneq H_{2}\supsetneq\cdots\subsetneq H_{2t}\supsetneq 0

where we will have the convention that Hj=IsjH_{j}=I_{s_{j}} for all 1j2t1\leqslant j\leqslant 2t. We also declare that s0=0s_{0}=0 and s2t+1=m+1s_{2t+1}=m+1.

Since this is a path in the graph GA2(R)GA_{2}(R), then successive ideals must have maximal conductor. We will say that Mk=(Ik+1:Ik)M_{k}=(I_{k+1}:I_{k}) if s2aks2a+1s_{2a}\leqslant k\leqslant s_{2a+1} for 0at0\leqslant a\leqslant t.

We first note that

H1=Is1𝔐M0M1Ms11H_{1}=I_{s_{1}}\supseteq\mathfrak{M}M_{0}M_{1}\cdots M_{s_{1}-1}

and since H1H2H_{1}\subseteq H_{2}, we have that

H2𝔐M0M1Ms11.H_{2}\supseteq\mathfrak{M}M_{0}M_{1}\cdots M_{s_{1}-1}.

In a similar fashion, we note that

H3=Is3H2Ms2Ms2+1Ms31𝔐M0M1Ms11Ms2Ms2+1Ms31.H_{3}=I_{s_{3}}\supseteq H_{2}M_{s_{2}}M_{s_{2}+1}\cdots M_{s_{3}-1}\supseteq\mathfrak{M}M_{0}M_{1}\cdots M_{s_{1}-1}M_{s_{2}}M_{s_{2}+1}\cdots M_{s_{3}-1}.

Inductively we obtain

H2i+1𝔐M0M1Ms11Ms2Ms2+1Ms31Ms2iMs2i+1Ms2i+11.H_{2i+1}\supseteq\mathfrak{M}M_{0}M_{1}\cdots M_{s_{1}-1}M_{s_{2}}M_{s_{2}+1}\cdots M_{s_{3}-1}\cdots M_{s_{2i}}M_{s_{2i}+1}\cdots M_{s_{2i+1}-1}.

In particular we obtain

0=𝔐M0M1Ms11Ms2Ms2+1Ms31Ms2tMs2t+1Mm0=\mathfrak{M}M_{0}M_{1}\cdots M_{s_{1}-1}M_{s_{2}}M_{s_{2}+1}\cdots M_{s_{3}-1}\cdots M_{s_{2t}}M_{s_{2t}+1}\cdots M_{m}

and hence there is a collection of maximal ideals with product equal to (0)(0).

For the implication (3)(1)(3)\Longrightarrow(1), we will assume that there is a collection of maximal ideals 𝔐i, 1in\mathfrak{M}_{i},\ 1\leqslant i\leqslant n such that 𝔐1𝔐2𝔐n=0\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0. To show that GA1(R)GA_{1}(R) is connected, it suffices to show that if IRI\subset R is an arbitrary ideal, then there is a finite path to the zero ideal. But note that

II𝔐1I𝔐1𝔐2I𝔐1𝔐2𝔐n=0I\supseteq I\mathfrak{M}_{1}\supseteq I\mathfrak{M}_{1}\mathfrak{M}_{2}\supseteq\cdots\supseteq I\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0

is such a path of length no more than nn. ∎

Corollary 3.17.

If GAi(R), 1i2GA_{i}(R),\ 1\leqslant i\leqslant 2 is connected then diam(GAi(R))2n\text{diam}(GA_{i}(R))\leqslant 2n where nn is the smallest positive integer for which there is a collection of maximal ideals 𝔐i,1in\mathfrak{M}_{i},1\leqslant i\leqslant n for which 𝔐1𝔐2𝔐n=0\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0.

Proof.

The fact that diam(GA1(R))2n\text{diam}(GA_{1}(R))\leqslant 2n is immediate from the proof of the implication (3)(1)(3)\Longrightarrow(1) in Theorem 3.16. The fact that diam(GA2(R))2n\text{diam}(GA_{2}(R))\leqslant 2n follows from the fact that GA1(R)GA_{1}(R) is a subgraph of GA2(R)GA_{2}(R) (see Proposition 3.15). ∎

Corollary 3.18.

If GAi(R)GA_{i}(R) is connected for some 0i20\leqslant i\leqslant 2 then RR is semiquasilocal and 00-dimensional.

Proof.

By Theorem 3.13, GA0(R)GA_{0}(R) is connected if and only if RR is Artinian, and hence RR is 00-dimensional, and, in this case, semilocal. If GA1(R)GA_{1}(R) or GA2(R)GA_{2}(R) is connected then Theorem 3.16 gives that 𝔐1𝔐2𝔐n=0\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0 for (not necessarily distinct) ideals {𝔐1,𝔐2,,𝔐n}\{\mathfrak{M}_{1},\mathfrak{M}_{2},\cdots,\mathfrak{M}_{n}\}. If 𝔐\mathfrak{M} is an arbitrary maximal ideal, then 𝔐𝔐1𝔐2𝔐n\mathfrak{M}\supseteq\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n} and hence 𝔐=𝔐k\mathfrak{M}=\mathfrak{M}_{k} for some 1kn1\leqslant k\leqslant n which shows that the list of ideals {𝔐1,𝔐2,,𝔐n}\{\mathfrak{M}_{1},\mathfrak{M}_{2},\cdots,\mathfrak{M}_{n}\} contains MaxSpec(R)\text{MaxSpec}(R). Hence RR is semiquasilocal.

To see that RR is 00-dimensional, we appeal once again to the fact that

𝔐1𝔐2𝔐n=0.\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0.

Recalling that this collection of maximal ideals contains MaxSpec(R)\text{MaxSpec}(R), we suppose that we can find a prime ideal 𝔓\mathfrak{P} such that 𝔐k𝔓\mathfrak{M}_{k}\supsetneq\mathfrak{P}. Since 𝔓𝔐1𝔐2𝔐n=0\mathfrak{P}\supseteq\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0, we must have that 𝔓𝔐i\mathfrak{P}\supseteq\mathfrak{M}_{i} for some 1in1\leqslant i\leqslant n. Hence 𝔐i𝔐k\mathfrak{M}_{i}\subsetneq\mathfrak{M}_{k} which is our desired contradiction. ∎

As a companion to Proposition 3.15, we present the following corollary.

Corollary 3.19.

If GA0(R)GA_{0}(R) is connected, then both GA1(R)GA_{1}(R) and GA2(R)GA_{2}(R) are connected.

Proof.

Combine the results from Proposition 3.15 and Theorem 3.16. ∎

As a final observation, we consider the following.

Corollary 3.20.

Let RR be a commutative ring with identity. We consider the following conditions:

  1. (1)

    GA0(R)GA_{0}(R) is connected.

  2. (2)

    GA1(R)GA_{1}(R) is connected.

  3. (3)

    GA2(R)GA_{2}(R) is connected.

If RR is Noetherian, the above three conditions are equivalent.

Proof.

We know that (2) and (3) are equivalent under any conditions and that condition (1) implies (2) and (3). Suppose that RR is Noetherian and any one of the conditions hold. By Corollary 3.18, RR is 00-dimensional. So if we add “Noetherian” as a hypothesis, then RR is Artinian and all three conditions hold. ∎

We conclude this section with an example of a ring RR for which GA0(R)GA_{0}(R) is not connected, but GAi(R)GA_{i}(R) is connected for i=1,2i=1,2.

Example 3.21.

Consider the ring

R:=𝔽[x1,x2,x3,]/({xixj}i,j1).R:=\mathbb{F}[x_{1},x_{2},x_{3},\cdots]/(\{x_{i}x_{j}\}_{i,j\geqslant 1}).

This ring is not Noetherian (and hence not Artinian) so GA0(R)GA_{0}(R) is not connected. On the other hand, GA1(R)GA_{1}(R) and GA2(R)GA_{2}(R) are connected. To see this, note that the unique maximal ideal of 𝔐R\mathfrak{M}\subseteq R (generated by the image of the elements xix_{i}) has the property that 𝔐2=0\mathfrak{M}^{2}=0. Hence every nonzero ideal of RR is connected by an edge to (0)(0) in both GA1(R)GA_{1}(R) and GA2(R)GA_{2}(R).

4. Class Structure: GClint(R)GCl_{int}(R) and GCl(R)GCl(R)

For the graphs GCl(R)GCl(R) and GClint(R)GCl_{int}(R) we assume that RR is an integral domain with quotient field KK unless specified otherwise.

It should be noted that in the case that RR is an integral domain, GClint(R)GCl_{int}(R) is a variant on the so-called divisor graph of an integral domain studied in [8], where the ideals II and JJ are assumed to be principal and possess an edge between them if I=JaI=Ja where aRa\in R is irreducible.

For this section, it will also be useful to keep in mind that GClint(R)GCl_{int}(R) is a subgraph of GCl(R)GCl(R).

Theorem 4.1.

The following conditions are equivalent.

  1. (1)

    GCl(R)GCl(R) is connected.

  2. (2)

    GCl(R)GCl(R) is complete.

  3. (3)

    RR is a PID

Proof.

For this proof, we discard the case where RR is a field as all of the conditions are satisfied vacuously. Since any complete graph is connected, (2)(1)(2)\Longrightarrow(1) is immediate. For the implication (1)(3)(1)\Longrightarrow(3), we let IRI\subset R be an arbitrary ideal and RR the unit ideal. Since GCl(R)GCl(R) is connected, there is a sequence of ideals connecting RR and II:

R:=J0-J1-J2--Jn1-I:=Jn.R:=J_{0}\relbar J_{1}\relbar J_{2}\relbar\cdots\relbar J_{n-1}\relbar I:=J_{n}.

Since the edges above are in the graph GCl(R)GCl(R), we must have, for all 1in1\leqslant i\leqslant n, nonzero kiKk_{i}\in K such that Ji1=kiJiJ_{i-1}=k_{i}J_{i}. Note that R=k1k2knIR=k_{1}k_{2}\cdots k_{n}I and hence, II is principal.

Finally for the implication (3)(2)(3)\Longrightarrow(2), we let I=aRI=aR and J=bRJ=bR be two arbitrary ideals of RR (with a,b0a,b\neq 0). Note that I=abJI=\frac{a}{b}J and hence GCl(R)GCl(R) is complete. ∎

Theorem 4.2.

GClint(R)GCl_{int}(R) is complete if and only if RR is a Noetherian valuation domain.

Proof.

In this proof, we will ignore the simple case where RR is a field. For the forward implication, we will assume that GClint(R)GCl_{int}(R) is complete. As GClint(R)GCl_{int}(R) is a subgraph of GCl(R)GCl(R), GCl(R)GCl(R) must also be complete. By Theorem 4.1, RR must be a PID. It now suffices to show that RR is local. To this end, suppose that 𝔐1𝔐2\mathfrak{M}_{1}\neq\mathfrak{M}_{2} are maximal ideals. Without loss of generality, there is a nonzero aRa\in R such that 𝔐1=a𝔐2\mathfrak{M}_{1}=a\mathfrak{M}_{2}. If aa is a unit then 𝔐1=𝔐2\mathfrak{M}_{1}=\mathfrak{M}_{2} and if aa is a nonunit then 𝔐1=a𝔐2𝔐2\mathfrak{M}_{1}=a\mathfrak{M}_{2}\subsetneq\mathfrak{M}_{2}; either gives a contradiction. Hence RR is a local PID and hence a Noetherian valuation domain.

On the other hand, if RR is a Noetherian valuation domain then any two nonzero proper ideals are of the form (πn)(\pi^{n}) and (πm)(\pi^{m}) where π\pi is a generator of the maximal ideal and n,m1n,m\geqslant 1. If we say (without loss of generality) that nmn\leqslant m then πmn(πn)=(πm)\pi^{m-n}(\pi^{n})=(\pi^{m}) and hence GClint(R)GCl_{int}(R) is complete. ∎

Theorem 4.3.

GClint(R)GCl_{int}(R) is connected if and only if RR is a PID. In this case, diam(GClint(R))2\text{diam}(GCl_{int}(R))\leqslant 2, and diam(GClint(R))=1\text{diam}(GCl_{int}(R))=1 if and only if RR is local.

Proof.

As GClint(R)GCl_{int}(R) is a subgraph of GCl(R)GCl(R), the fact that GClint(R)GCl_{int}(R) is connected implies that GCl(R)GCl(R) is connected. Hence, by Theorem 4.1, RR must be a PID.

On the other hand, if RR is a PID and I=aRI=aR and J=bRJ=bR (a,b0a,b\neq 0) are arbitrary ideals, then we can connect II and JJ as follows:

I=aRabRbR=J.I=aR\leftrightarrow abR\leftrightarrow bR=J.

The above demonstrates the veracity of the remark that diam(GClint(R))2\text{diam}(GCl_{int}(R))\leqslant 2. The fact that diam(GClint(R))=1\text{diam}(GCl_{int}(R))=1 precisely when RR is local follows from Theorem 4.2. ∎

Theorem 4.4.

If RR is a Dedekind domain with quotient field KK, then the connected components of the graphs GCl(R)GCl(R) and GClint(R)GCl_{int}(R) are in one to one correspondence with the elements of the class group Cl(R)\text{Cl}(R). Each connected component of GCl(R)GCl(R) is complete and each connected component of GClint(R)GCl_{int}(R) has diameter no more than 2 and the connected components of GClint(R)GCl_{int}(R) are complete if and only if RR is a local PID.

Proof.

If RR is a Dedekind domain with quotient field KK, then two ideals, II and JJ, are in the same class of Cl(R)\text{Cl}(R) if and only if I=JkI=Jk for some nonzero kKk\in K. Hence each connected component of GCl(R)GCl(R) is complete and these components are in one to one correspondence with the elements of Cl(R)\text{Cl}(R).

For the GClint(R)GCl_{int}(R) case, we first note that if there is a path from II to JJ then there must be some nonzero kKk\in K such that I=JkI=Jk, so it remains to show that if II and JJ are in the same ideal class, then there is a path connecting them. To this end, we note that if I=JkI=Jk for some nonzero kKk\in K, we can write I=abJI=\frac{a}{b}J and in a similar fashion as before, we can connect II and JJ via

IbI=aJJ,I\leftrightarrow bI=aJ\leftrightarrow J,

and hence there is a path of length no more than 2 connecting II and JJ.

For the final statement, we first suppose that RR is a local PID (and hence a Noetherian valuation domain) with uniformizer π\pi. If II and JJ are in the same class of Cl(R)\text{Cl}(R) then I=kJI=kJ for some nonzero kKk\in K. Since KK is the quotient field of a Noetherian valuation domain, we can write k=πnk=\pi^{n} with nn\in\mathbb{Z} up to a unit in RR. If n0n\geqslant 0 then I=JI=J or there is an edge between them. If n<0n<0 then J=πnIJ=\pi^{-n}I and again there is an edge between II and JJ.

On the other hand, suppose that some connected component of GClint(R)GCl_{int}(R) is complete and that there are two maximal ideals 𝔐1,𝔐2R\mathfrak{M}_{1},\mathfrak{M}_{2}\subset R. Select m1𝔐1𝔐2m_{1}\in\mathfrak{M}_{1}\setminus\mathfrak{M}_{2} and m2𝔐2𝔐1m_{2}\in\mathfrak{M}_{2}\setminus\mathfrak{M}_{1}. If II is in a complete connected component of GClint(R)GCl_{int}(R), then there is a path from m1Im_{1}I to m2Im_{2}I and hence there must be an edge between them. Hence we have m1|m2m_{1}|m_{2} or m2|m1m_{2}|m_{1} and in either case, we obtain a contradiction. We conclude that RR is local and hence a PID (Noetherian valuation domain). ∎

In the spirit of these results, we further restrict the set of vertices to make a more general observation. We define GClinv(R)GCl_{inv}(R) to be the subgraph of GCl(R)GCl(R) with the vertex set restricted to the set of invertible ideals. Recall that if RR is a domain with quotient field KK, then RR is seminormal if for all ωK\omega\in K, ω2,ω3R\omega^{2},\omega^{3}\in R implies that ωR\omega\in R (see [19] for a good reference on this topic). The next result shows that if RR is a seminormal domain then the number of connected components in GClinv(R)GCl_{inv}(R) is stable for polynomial extensions.

Theorem 4.5.

If RR is a seminormal domain, then the number of connected components of GClinv(R)GCl_{inv}(R) is equal to the number of connected components of
GClinv(R[x1,x2,,xn])GCl_{inv}(R[x_{1},x_{2},\cdots,x_{n}]) for all n1n\geqslant 1.

Proof.

This result follows from the observations that in the integral domain case Pic(R)\text{Pic}(R) is isomorphic to Cl(R)\text{Cl}(R) and Pic(R)Pic(R[x])\text{Pic}(R)\cong\text{Pic}(R[x]) ([11]) and the fact that a polynomial extension of a seminormal domain is seminormal ([19]). With these results in hand, the fact that we are restricting the vertex set to the set of invertible ideals makes the rest of the proof almost identical to the proof of Theorem 4.4. ∎

5. GZ(R)GZ(R) and the classical zero-divisor graph

In this section, our attention will be devoted to the graph GZ(R)GZ(R) and some of its variants. The reason for excluding the zero ideal is precisely the same as the reason that this exception was made in [2]: namely because the use of the zero ideal gives extra structure to this graph with no useful new information. It is easy to see that for any commutative ring with identity (even an integral domain) that the graph GZ(R)GZ(R) is connected with diameter no more than 22 if we allow use of the zero ideal. Indeed, if II and JJ are arbitrary ideals, then I(0)JI\leftrightarrow(0)\leftrightarrow J is a path connecting them. So if RR is an (infinite) integral domain, the graph GZ(R)GZ(R) would be an infinite star graph with the zero ideal at the center. These extra connections muddy the waters and give no useful insights for our current purposes.

The first variant that we will consider is the graph GZ(R)GZ^{*}(R). For this graph, the set of vertices is the collection of nonzero ideals IRI\subset R such that there is a nonzero ideal JRJ\subset R such that IJ=0IJ=0. As before, we say that II and JJ have an edge between them if IJ=0IJ=0. Of course, if RR is a domain, this produces the empty graph, so in this situation, RR must have nontrivial zero divisors for this graph to be of any interest.

We begin with a well-known lemma concerning annihilators that we record here for completeness with proof omitted.

Lemma 5.1.

Let IJI\subseteq J be ideals in RR.

  1. (1)

    Ann(J)Ann(I)\text{Ann}(J)\subseteq\text{Ann}(I).

  2. (2)

    IAnn(Ann(I))I\subseteq\text{Ann}(\text{Ann}(I)).

  3. (3)

    Ann(I)=Ann(Ann(Ann(I)))\text{Ann}(I)=\text{Ann}(\text{Ann}(\text{Ann}(I))).

The next theorem reminiscent of Theorem 1.1 from [2] couched in an ideal-theoretic setting. This theorem can also be found in [7] and may be considered a consequence of [10, Theorem 11]. This will be used to leverage some further insights and to provide an alternate proof of the first statement in [2, Theorem 2.3].

Theorem 5.2.

Let RR be a commutative ring with 1, then GZ(R)GZ^{*}(R) is connected and has diameter no more than 3.

Proof.

Let II and JJ be distinct nonzero ideals of RR such that there are ideals 0I1,J1R0\neq I_{1},J_{1}\subset R with II1=0=JJ1II_{1}=0=JJ_{1}. As before, we use the notation “XYX\leftrightarrow Y” to mean that X=YX=Y or that there is an edge between XX and YY.

We first note that if JAnn(I)0J\text{Ann}(I)\neq 0 then we have the path

IJAnn(I)Ann(J)JI\leftrightarrow J\text{Ann}(I)\leftrightarrow\text{Ann}(J)\leftrightarrow J

and similarly, if IAnn(J)0I\text{Ann}(J)\neq 0 we have the path

IAnn(I)IAnn(J)JI\leftrightarrow\text{Ann}(I)\leftrightarrow I\text{Ann}(J)\leftrightarrow J

So the only case to consider is the case where JAnn(I)=IAnn(J)=0J\text{Ann}(I)=I\text{Ann}(J)=0. Note that this implies that Ann(I)Ann(J)\text{Ann}(I)\subseteq\text{Ann}(J) and Ann(J)Ann(I)\text{Ann}(J)\subseteq\text{Ann}(I). Since Ann(I)=Ann(J)\text{Ann}(I)=\text{Ann}(J) we have the path

IAnn(I)J.I\leftrightarrow\text{Ann}(I)\leftrightarrow J.

We now recover the famous result of D. F. Anderson and P. Livingston.

Corollary 5.3 (D. F. Anderson and P. Livingston,[2]).

If Z(R)Z(R) is the zero divisor graph of a commutative ring with identity, then Z(R)Z(R) is connected with diameter no more than 3.

Proof.

In the proof of Theorem 5.2, we now restrict to the case where II and JJ are principal (and note that principal subideals can be chosen from within each respective annihilator). This shows that the graph of principal ideals that are generated by zero-divisors satisfy the conclusion of the corollary. To see that the slightly different statement formulated by Anderson and Livingston holds, just note that if a,bRa,b\in R are distinct nonzero zero-divisors such that (a)(b)(a)\neq(b), then there is a path among nonzero principal ideals of the form

(a)(x)(y)(b)(a)\leftrightarrow(x)\leftrightarrow(y)\leftrightarrow(b)

or

(a)(x)(b).(a)\leftrightarrow(x)\leftrightarrow(b).

This yields the path axyba\leftrightarrow x\leftrightarrow y\leftrightarrow b or axba\leftrightarrow x\leftrightarrow b. In the case (a)=(b)(a)=(b), we note that there is an element 0c0\neq c such that ac=0=bcac=0=bc giving the path acba\leftrightarrow c\leftrightarrow b unless a=ca=c (without loss of generality). But in this case, a2=0=aba^{2}=0=ab and we are done. ∎

We now return to GZ(R)GZ(R). Note that if RR contains a regular element xx then the ideal (x)(x) is an isolated vertex (and so, in general, one does not expect this graph to be connected). We highlight a case where connection is forced.

Theorem 5.4.

Let RR be a commutative ring with identity that is not a field. If GA1(R)GA_{1}(R) or GA2(R)GA_{2}(R) is connected, then so is GZ(R)GZ(R). Additionally diam(GZ(R))3\text{diam}(GZ(R))\leqslant 3.

Proof.

By Theorem 3.16 there is a collection of not necessarily distinct maximal ideals {𝔐1,𝔐2,,𝔐n}\{\mathfrak{M}_{1},\mathfrak{M}_{2},\cdots,\mathfrak{M}_{n}\} such that 𝔐1𝔐2𝔐n=0,n2\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}=0,\ n\geqslant 2. We can also assume that no proper subproduct of these listed ideals is zero.

If I,JRI,J\subset R then there is a maximal ideal 𝔐\mathfrak{M} containing II. Since 𝔐I0=𝔐1𝔐2𝔐n\mathfrak{M}\supseteq I\supset 0=\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{n}, 𝔐\mathfrak{M} must be 𝔐i\mathfrak{M}_{i} for some 1in1\leqslant i\leqslant n. We will assume without loss of generality that I𝔐1I\subseteq\mathfrak{M}_{1} and J𝔐k, 1knJ\subseteq\mathfrak{M}_{k},\ 1\leqslant k\leqslant n. To see that GZ(R)GZ(R) is connected with diameter no more than three consider the path for the case where k1k\neq 1:

I𝔐2𝔐3𝔐n𝔐1𝔐2𝔐k1𝔐k+1𝔐nJI\leftrightarrow\mathfrak{M}_{2}\mathfrak{M}_{3}\cdots\mathfrak{M}_{n}\leftrightarrow\mathfrak{M}_{1}\mathfrak{M}_{2}\cdots\mathfrak{M}_{k-1}\mathfrak{M}_{k+1}\cdots\mathfrak{M}_{n}\leftrightarrow J

In the case that k=1k=1, we modify our path as follows:

I𝔐2𝔐3𝔐nJ.I\leftrightarrow\mathfrak{M}_{2}\mathfrak{M}_{3}\cdots\mathfrak{M}_{n}\leftrightarrow J.

We conclude this section with an example to show that GZ(R)GZ(R) may be connected without GA1(R)GA_{1}(R), GA2(R)GA_{2}(R) being connected.

Example 5.5.

Let 𝔽\mathbb{F} be a field and consider first the domain R:=𝔽[x1,x2,,xn,,y]R:=\mathbb{F}[x_{1},x_{2},\cdots,x_{n},\cdots,y], let IRI\subset R be the ideal I:=({xi2,xiy,y2}i1)I:=(\{x_{i}^{2},x_{i}y,y^{2}\}_{i\geqslant 1}), and let T:=R/IT:=R/I (we abuse the notation by now thinking of the elements y,xiy,x_{i} as elements of TT). TT is quasilocal, with maximal ideal 𝔐\mathfrak{M}. It is easy to see that GZ(T)GZ(T) is connected. Indeed if I,JTI,J\subset T then we have the path I(y)JI\leftrightarrow(y)\leftrightarrow J. But there is no collection of maximal ideals with product 0 since for all k1, 0x1x2xk𝔐kk\geqslant 1,\ 0\neq x_{1}x_{2}\cdots x_{k}\in\mathfrak{M}^{k}.

6. Finite Containment: GF(R) and GP(R)GF(R)\text{ and }GP(R)

In this section we investigate graphs of rings where the edges are defined to highlight finite (or perhaps principal) generation of an ideal over a subideal that it contains. In these graphs, the vertices will be the set of proper ideals and edges between ideals will be defined by finite generation. Specifically, we will declare that the ideals II and JJ have an edge between them if IJI\subset J and JJ is finitely (resp. principally) generated over II; that is, J=(I,x1,x2,,xn)J=(I,x_{1},x_{2},\cdots,x_{n}) (resp. J=(I,x)J=(I,x)).

The algebraic motivation for this definition is an attempt to measure properties of MaxSpec(R)\text{MaxSpec}(R) in a condition that (under certain graphical constraints) mimics the Noetherian condition.

This first lemma is very easy and is presented to connect our definition of finite generation of one ideal over another to what was presented in Definition 2.3. We record it for completeness and omit the proof.

Lemma 6.1.

If IJRI\subseteq J\subset R be ideals of RR, then J/IJ/I is a finitely generated ideal of R/IR/I if and only if JJ is finitely generated over II.

Here are some preliminary containments of note.

Proposition 6.2.

GA0(R)GA_{0}(R) is a subgraph of GP(R)GP(R) and GP(R)GP(R) is a subgraph of GF(R)GF(R).

Proof.

Only the first statement needs proof. Suppose that IJRI\subsetneq J\subsetneq R are adjacent ideals (so there is an edge between them in GA0(R)GA_{0}(R)). If xJIx\in J\setminus I then J=(I,x)J=(I,x) by adjacency and hence II and JJ have an edge between them in GP(R)GP(R). ∎

We now present the quasilocal case, which we will see is an exceptional case for these graphs in the sense that they are not always connected. Perhaps surprisingly, outside the quasilocal case, GF(R)GF(R) and GP(R)GP(R) are always connected.

Theorem 6.3.

Let (R,𝔐)(R,\mathfrak{M}) be quasilocal. GF(R)GF(R) (resp. GP(R)GP(R)) is connected if and only if 𝔐\mathfrak{M} is finitely generated. Additionally, we have the following.

  1. (1)

    If GF(R)GF(R) is connected then diam(GF(R))2\text{diam}(GF(R))\leqslant 2 and GF(R)GF(R) is complete if and only if RR is a Noetherian chained ring of dimension no more than 11.

  2. (2)

    If GP(R)GP(R) is connected then diam(GP(R))2n\text{diam}(GP(R))\leqslant 2n where nn is the minimal number of generators required for 𝔐\mathfrak{M} and GP(R)GP(R) is complete if and only if RR is a PIR.

Proof.

We first remark that if RR is a field then the result hold trivially, so we will assume that RR is not a field. For the initial statement, we first suppose that 𝔐\mathfrak{M} is finitely generated. To show that GF(R)GF(R) is connected, we first note that since 𝔐\mathfrak{M} is finitely generated, it is certainly finitely generated over II. Hence if I,JRI,J\subset R are any two ideals of RR, then I𝔐JI\leftrightarrow\mathfrak{M}\leftrightarrow J is a path of length no more than 2 from II to JJ. For GP(R)GP(R) the proof is similar (with a possibly longer path).

Now we suppose that GF(R)GF(R) is connected. By assumption, there is a finite path from (0)(0) to 𝔐\mathfrak{M}. Since a finite extension of a finite extension is finite, we can assume that this path takes on the form

(0)I1J1I2J2In1Jn1In=𝔐.(0)\subset I_{1}\supset J_{1}\subset I_{2}\supset J_{2}\subset\cdots\subset I_{n-1}\supset J_{n-1}\subset I_{n}=\mathfrak{M}.

Note that I1=(a1,1,a1,2,,a1,t1)I_{1}=(a_{1,1},a_{1,2},\cdots,a_{1,t_{1}}) is finitely generated and for all 2mn2\leqslant m\leqslant n we can write Im=(Jm1,am,1,am,2,,am,tm)I_{m}=(J_{m-1},a_{m,1},a_{m,2},\cdots,a_{m,t_{m}}), and since Jm1Im1J_{m-1}\subset I_{m-1}, we have that Im(Im1,am,1,am,2,,am,tm)I_{m}\subseteq(I_{m-1},a_{m,1},a_{m,2},\cdots,a_{m,t_{m}})

In particular,

In=𝔐(I1,{ai,j}j=1ti)i=2nI_{n}=\mathfrak{M}\subseteq(I_{1},\{a_{i,j}\}_{j=1}^{t_{i}}{{}_{i=2}^{n}})

and since 𝔐\mathfrak{M} is maximal, equality holds. Hence 𝔐\mathfrak{M} is finitely generated. Since each finite extension is a finite sequence of principal extensions, this establishes the statement for GP(R)GP(R) as well.

For (1), we have already established that diam(GF(R))2\text{diam}(GF(R))\leqslant 2. If GF(R)GF(R) is complete, then any two ideals must be comparable and hence RR is chained. Now note that if II is an arbitrary ideal of RR then there is an edge between II and (0)(0) and so II must be finitely generated and so RR is Noetherian. Finally, we note that since RR is Noetherian and chained, its dimension must be no more than 11 (if RR has a height 11 nonmaximal prime ideal then there must be infinitely many and hence RR cannot be chained).

For the converse, note that if RR is Noetherian and chained then any two ideals are comparable and since they are finitely generated, there must be an edge between them.

For (2) we note that if 𝔐\mathfrak{M} is generated by {x1,x2,,xn}\{x_{1},x_{2},\cdots,x_{n}\} then as RR is quasilocal, there is a path from any ideal II to 𝔐\mathfrak{M} of length no more than nn. So given ideals I,JRI,J\subset R, there is a path from II to JJ bounded by twice the number of generators of 𝔐\mathfrak{M} and hence diam(GP(R))2n\text{diam}(GP(R))\leqslant 2n.

Now suppose that GP(R)GP(R) is complete and let IRI\subset R be an ideal. Since II and (0)(0) have an edge between them, II must be principal. Conversely if RR is a PIR and local then RR is chained. To see this, note that if I=(x)I=(x) and J=(y)J=(y) then (d)=(x,y)(d)=(x,y). If dx=xdx^{\prime}=x and dy=ydy^{\prime}=y, it is easy to see that (x,y)=R(x^{\prime},y^{\prime})=R and so (without loss of generality) xx^{\prime} is a unit and x|yx|y. Hence JIJ\subseteq I and so there is an edge between II and JJ. ∎

We now suppose that RR is not quasilocal. In this case, GF(R)GF(R) and GP(R)GP(R) are always connected, but as we will see, the diameter of GF(R)GF(R) reveals some subtleties concerning the structure of RR. With regard to diameter, we will focus on GF(R)GF(R) as the large number of steps sometimes required to make a path in GP(R)GP(R) clouds the issue a bit. On the other hand, we will make note of GP(R)GP(R) when prudent.

We begin with a definition and a useful lemma that will simplify matters.

Definition 6.4.

Let IRI\subset R be an ideal. We define MaxSpecI(R)\text{MaxSpec}_{I}(R) to be the collection of maximal ideals of RR containing II and following [12] we define the Jacobson radical of II, 𝔍(I)\mathfrak{J}(I), to be the intersection of all maximal ideals of RR containing II. If RR is a ring we use the notation 𝔍(R)\mathfrak{J}(R) to be the Jacobson radical of RR.

We remark that it is an easy exercise to verify that 𝔍(R/I)=𝔍(I)/I\mathfrak{J}(R/I)=\mathfrak{J}(I)/I and we will use this fact on a number of occasions.

In the following key lemma, we describe paths between ideals in GF(R)GF(R) and GP(R)GP(R). The bounds apply to both as principally generated ideals will be used in the proof.

Lemma 6.5.

Let RR be a commutative ring with identity and consider the graphs GF(R)GF(R) and GP(R)GP(R).

  1. (1)

    If the ideals II and JJ are comaximal, then there is a path of length 22 from II to JJ.

  2. (2)

    If II and JJ are not comaximal, but there is a maximal ideal MM that contains II but not JJ then there is a path of length no more than 33 from II to JJ.

  3. (3)

    If MaxSpecI(R)=MaxSpecJ(R)\text{MaxSpec}_{I}(R)=\text{MaxSpec}_{J}(R) and |MaxSpec(R)|>1|\text{MaxSpec}(R)|>1, then there is a path of length no more than 44 from II to JJ.

Proof.

For (1), we suppose that II and JJ are comaximal and find αI\alpha\in I and βJ\beta\in J such that α+β=1\alpha+\beta=1. So if xIx\in I (resp. xJx\in J) then the equation

xα+xβ=xx\alpha+x\beta=x

demonstrates that the ideal II (resp. JJ) is singly generated by α\alpha (resp. β\beta) over the ideal IJIJ. Hence in GF(R)GF(R) (as well as GP(R)GP(R)), we have the path IIJJI-IJ-J of length 2.

For (2), since MM does not contain JJ and is maximal, we can find mMm\in M, jJj\in J such that m+j=1m+j=1. Since IMI\subseteq M we have that (I,m)(I,m) is a proper ideal and is clearly principally generated (or less) over II. By (1) there is a path between (I,m)(I,m) and JJ of length 22 and hence there is a path of length no more than 33 from II to JJ.

For the last statement, we will assume that MaxSpecI(R)=MaxSpecJ(R)\text{MaxSpec}_{I}(R)=\text{MaxSpec}_{J}(R) and select MMaxSpecI(R)M\in\text{MaxSpec}_{I}(R). By hypothesis, there is another maximal ideal NN and we find mMm\in M and nNn\in N such that m+n=1m+n=1. In any case, we have the ideals II and (I,m)(I,m) are equal or have an edge between them. If JNJ\subseteq N then JJ and (J,n)(J,n) are equal or have an edge between them and since (I,m)(I,m) and (J,n)(J,n) are comaximal, we have our desired path of length no more than 4. On the other hand, if JJ is not contained in NN, then JJ and NN are comaximal (as are II and NN). From part (1) there is a path of length 2 from II to NN and a path of length 2 from NN to JJ and this completes the proof. ∎

The following theorem follows directly from Lemma 6.5.

Theorem 6.6.

Let RR be commutative with 1 with |MaxSpec(R)|>1|\text{MaxSpec}(R)|>1. Then GF(R)GF(R) and GP(R)GP(R) are connected and diam(R)4\text{diam}(R)\leqslant 4.

We now focus on GF(R)GF(R) in the case that |MaxSpec(R)|>1|\text{MaxSpec}(R)|>1 and examine necessary and sufficient conditions for the diameter to be of prescribed sizes.

Theorem 6.7.

Let RR be a commutative ring with 1 with |MaxSpec(R)|>1|\text{MaxSpec}(R)|>1.

  1. (1)

    diam(GF(R))=2\text{diam}(GF(R))=2 if and only if every maximal ideal of RR is finitely generated.

  2. (2)

    diam(GF(R))3\text{diam}(GF(R))\leqslant 3 if and only if given I,JRI,J\subset R with 𝔍(I)=𝔍(J)\mathfrak{J}(I)=\mathfrak{J}(J) then there is a proper ideal KK that is finitely generated over both II and JJ.

Proof.

If diam(GF(R))=2\text{diam}(GF(R))=2, then given any maximal ideal 𝔐\mathfrak{M}, there is a path of length 11 or 22 between 𝔐\mathfrak{M} and (0)(0). Whether the path is of the form (0)-𝔐(0)\relbar\mathfrak{M} or (0)-I-𝔐(0)\relbar I\relbar\mathfrak{M}, 𝔐\mathfrak{M} is finitely generated.

On the other hand, if I,JRI,J\subset R are ideals, then there is a path of length 22 between them if II and JJ are comaximal by Lemma 6.5. If II and JJ are both contained in the maximal ideal 𝔐\mathfrak{M} then I𝔐JI\leftrightarrow\mathfrak{M}\leftrightarrow J is a path of length no more than 22 between them. Since |MaxSpec(R)|>1|\text{MaxSpec}(R)|>1, we have that diam(GF(R))\text{diam}(GF(R)) is precisely 22. This establishes the first statement.

For (2), we first suppose that diam(GF(R))3\text{diam}(GF(R))\leqslant 3. Let I,JRI,J\subset R be ideals with 𝔍(I)=𝔍(J)\mathfrak{J}(I)=\mathfrak{J}(J) and consider cases. If there is an edge between II and JJ, then the larger ideal is finitely generated over both.

If there is a path of length 22 between II and JJ, say IAJI-A-J and I,JAI,J\subset A then AA is finitely generated over both II and JJ. If on the other hand, AI,JA\subset I,J, then I=(A,x1,x2,,xn)I=(A,x_{1},x_{2},\cdots,x_{n}) and J=(A,y1,y2,,ym)J=(A,y_{1},y_{2},\cdots,y_{m}). Now note that I+J=(I,y1,y2,,ym)=(J,x1,x2,,xn)I+J=(I,y_{1},y_{2},\cdots,y_{m})=(J,x_{1},x_{2},\cdots,x_{n}) is finitely generated over both II and JJ and note that since 𝔍(I)=𝔍(J)\mathfrak{J}(I)=\mathfrak{J}(J), I+JI+J is proper.

Finally, we suppose that there is a path of length 33 between II and JJ of the form IABJI-A-B-J. Without loss of generality we will assume that we have the containments IABJI\supset A\subset B\supset J. In a similar fashion to the previous argument, we write I=(A,x1,x2,,xn)I=(A,x_{1},x_{2},\cdots,x_{n}) and B=(A,z1,z2,,zt)=(J,y1,y2,,ym)B=(A,z_{1},z_{2},\cdots,z_{t})=(J,y_{1},y_{2},\cdots,y_{m}). Note that I+BI+B is a proper ideal as 𝔍(I)=𝔍(J)𝔍(B)\mathfrak{J}(I)=\mathfrak{J}(J)\subseteq\mathfrak{J}(B). So the ideal I+B=(I,z1,z2,,zt)=(J,y1,y2,,ym,x1,x2,,xn)I+B=(I,z_{1},z_{2},\cdots,z_{t})=(J,y_{1},y_{2},\cdots,y_{m},x_{1},x_{2},\cdots,x_{n}) is finitely generated over both II and JJ.

Conversely, note that Lemma 6.5 shows that the only unresolved case is the case in which I,JRI,J\subset R with 𝔍(I)=𝔍(J)\mathfrak{J}(I)=\mathfrak{J}(J). But the existence of KK shows that there is a path between II and JJ of length no more than 22. Hence diam(GF(R))3\text{diam}(GF(R))\leqslant 3. ∎

Corollary 6.8.

Let RR be commutative with identity. Then diam(GF(R))3\text{diam}(GF(R))\leqslant 3 if and only if for all IRI\subset R, 𝔍(R/I)\mathfrak{J}(R/I) is contained in a finitely generated proper ideal.

Proof.

Note first that Theorem 6.3 shows that this result holds in the quasilocal case, and so we will assume that |MaxSpec(R)|>1|\text{MaxSpec}(R)|>1.

We suppose first that diam(GF(R))3\text{diam}(GF(R))\leqslant 3 and select an arbitrary IRI\subset R. Theorem 6.7 assures us that there is an ideal BRB\subset R with BB finitely generated over both II and 𝔍(I)\mathfrak{J}(I) and so B/IB/I is a finitely generated ideal of R/IR/I that contains 𝔍(I)/I=𝔍(R/I)\mathfrak{J}(I)/I=\mathfrak{J}(R/I).

Conversely, suppose that for all IRI\subset R we have that 𝔍(R/I)\mathfrak{J}(R/I) is contained in a finitely generated ideal and select ideals I,JRI,J\subset R with 𝔍(I)=𝔍(J)\mathfrak{J}(I)=\mathfrak{J}(J). By assumption 𝔍(R/IJ)=𝔍(IJ)/IJ\mathfrak{J}(R/IJ)=\mathfrak{J}(IJ)/IJ is contained in the finitely generated ideal K/IJK/IJ. Note that as MaxSpecI(R)=MaxSpecJ(R)\text{MaxSpec}_{I}(R)=\text{MaxSpec}_{J}(R), 𝔍(IJ)=𝔍(J)=𝔍(I)\mathfrak{J}(IJ)=\mathfrak{J}(J)=\mathfrak{J}(I) and so we have that K/IJK/IJ contains 𝔍(I)/IJ\mathfrak{J}(I)/IJ which in turn contains I/IJI/IJ. Hence K/I(K/IJ)/(I/IJ)K/I\cong(K/IJ)/(I/IJ) is finitely generated. A similar proof establishes that K/JK/J is finitely generated, and so KK is finitely generated over both II and JJ. By Theorem 6.7 we are done. ∎

Example 6.9.

Let VV be a 11-dimensional nondiscrete valuation domain with maximal ideal 𝔐\mathfrak{M} and consider the ring V[x]V[x]. Note that VV[x]/(x)V\cong V[x]/(x) has Jacobson radical 𝔐\mathfrak{M} and so by Corollary 6.8, GF(V[x])GF(V[x]) has diameter 44.

Example 6.10.

Consider the ring R:=iΓ𝔽2R:=\prod_{i\in\Gamma}\mathbb{F}_{2} where Γ\Gamma is a nonempty index set with at least two elements (the graph GF(R)GF(R) is a single vertex in the degenerate case that Γ\Gamma consists of a single element). Note first that any element of RR is idempotent and so this is true of any homomorphic image of RR. So if TT is a homomorphic image of RR and eTe\in T then e(1e)=0e(1-e)=0. So if e𝔍(T)e\in\mathfrak{J}(T) then 1e1-e is a unit in TT which implies that e=0e=0. Since 𝔍(T)=0\mathfrak{J}(T)=0 for all homomorphic images of TT of RR, we have that diam(GF(R))=3\text{diam}(GF(R))=3 if Γ\Gamma is infinite and diam(GF(R))=2\text{diam}(GF(R))=2 if Γ\Gamma is finite with at least two elements.

We now wish to show that the family of almost Dedekind domains that are not Dedekind produce examples of domains where diam(GF(R))=3\text{diam}(GF(R))=3 in abundance. We first require a couple of preliminary results.

Lemma 6.11.

If RR is a commutative ring with 1, SRS\subseteq R a multiplicative set, and IRI\subset R an ideal with the property that for all sSs\in S, there is an xsIx_{s}\in I and sSs^{\prime}\in S such that ss+xs=1s^{\prime}s+x_{s}=1, then R/IRS/ISR/I\cong R_{S}/I_{S}.

Proof.

We define ϕ:RRS/IS\phi:R\longrightarrow R_{S}/I_{S} given by ϕ(z)=z1+IS\phi(z)=\frac{z}{1}+I_{S}. We suppose that rRr\in R and sSs\in S. By hypothesis, we can find xsI,sSx_{s}\in I,s^{\prime}\in S with ss+xs=1s^{\prime}s+x_{s}=1. So in RSR_{S} we have s+xss=1ss^{\prime}+\frac{x_{s}}{s}=\frac{1}{s} and hence rs+rxss=rsrs^{\prime}+\frac{rx_{s}}{s}=\frac{r}{s}.

We now observe that ϕ(rs)=rs1+IS=rsrxss+IS=rs+IS\phi(rs^{\prime})=\frac{rs^{\prime}}{1}+I_{S}=\frac{r}{s}-\frac{rx_{s}}{s}+I_{S}=\frac{r}{s}+I_{S} and so ϕ\phi is onto.

It is easy to see that Iker(ϕ)I\subseteq\text{ker}(\phi). For the other containment, note that if zker(ϕ)z\in\text{ker}(\phi), then z1=αs\frac{z}{1}=\frac{\alpha}{s} for some αI\alpha\in I and sSs\in S. Hence there exists tSt\in S such that t(zsα)=0t(zs-\alpha)=0. Using the assumed property, there is a tSt^{\prime}\in S and βI\beta\in I such that tt+β=1tt^{\prime}+\beta=1. From this we obtain that (1β)(zsα)=0(1-\beta)(zs-\alpha)=0 and so zsIzs\in I. Once again, we use the fact that there is sSs^{\prime}\in S and γI\gamma\in I such that ss+γ=1ss^{\prime}+\gamma=1 and we now obtain that z(1γ)Iz(1-\gamma)\in I and hence zIz\in I. By the First Isomorphism Theorem, we have that R/IRS/ISR/I\cong R_{S}/I_{S}. ∎

Proposition 6.12.

Let IRI\subset R be an ideal and {Mi}iΓ\{M_{i}\}_{i\in\Gamma} be the collection of maximal ideals of RR that contain II. If S=(iΓMi)cS=(\bigcup_{i\in\Gamma}M_{i})^{c}, then R/IRS/ISR/I\cong R_{S}/I_{S}.

Proof.

Suppose that sSs\in S and note that by the definition of SS, (s,I)=R(s,I)=R. Hence we can find xsI,sRx_{s}\in I,s^{\prime}\in R with ss+xs=1s^{\prime}s+x_{s}=1. Observe that ss^{\prime} cannot be in any maximal ideal containing II and so must be in SS; we now appeal to Lemma 6.11. ∎

Proposition 6.13.

If RR be an almost Dedekind domain that is not Dedekind with finitely many maximal ideals that are not finitely generated, then diam(GF(R))=3\text{diam}(GF(R))=3.

We remark that almost Dedekind domains meeting the above conditions are commonplace. In particular, consider the sequence domains defined in [16].

Proof.

For the first case, we suppose that II is contained in only finitely many maximal ideals and M1,M2,,MnM_{1},M_{2},\cdots,M_{n} be the maximal ideals containing II. If S=(i=1nMi)cS=(\bigcup_{i=1}^{n}M_{i})^{c}, then Proposition 6.12 shows that R/I=RS/ISR/I=R_{S}/I_{S}, but as RSR_{S} is an almost Dedekind domain with only finitely many maximal ideals, it is Dedekind and hence 𝔍(R/I)𝔍(RS/IS)\mathfrak{J}(R/I)\cong\mathfrak{J}(R_{S}/I_{S}) is finitely generated. Hence by Corollary 6.8, diam(GF(R))3\text{diam}(GF(R))\leqslant 3.

Now suppose that II is contained in infinitely many maximal ideals. Note that since there are only finitely many maximal ideals that are not finitely generated, there must be infinitely many finitely generated primes (if not then RR is semiquasilocal and hence Dedekind). So if MM is a finitely generated maximal ideal containing II, then M/IM/I is a finitely generated ideal of R/IR/I containing 𝔍(R/I)=𝔍(I)/I\mathfrak{J}(R/I)=\mathfrak{J}(I)/I and so again diam(GF(R))3\text{diam}(GF(R))\leqslant 3 by Corollary 6.8.

Equality follows from the existence of maximal ideals that are not finitely generated and Theorem 6.7. ∎

We also distinguish behavior in the Noetherian case; we ignore the case of a field as the graph in this case is a single vertex.

Theorem 6.14.

Let RR be a commutative ring with identity that is not a field. The following conditions are equivalent.

  1. (1)

    RR is Noetherian.

  2. (2)

    The radius of GF(R)GF(R) is equal to 11, and (0)(0) is a center of GF(R)GF(R).

  3. (3)

    diam(GF(R))2\text{diam}(GF(R))\leqslant 2 and if I,JRI,J\subset R then either II and JJ have an edge between them or there is a minimal path between them passing through (0)(0).

Proof.

For (1)(2)(1)\Longrightarrow(2) we assume that RR is Noetherian. In this case there is an edge between (0)(0) and an arbitrary nonzero ideal. So (0)(0) is a center and the radius of GF(R)GF(R) is precisely 11.

For (2)(3)(2)\Longrightarrow(3), since (0)(0) is a center and has an edge with an arbitrary ideal, every ideal is finitely generated. In particular, each maximal ideal is finitely generated and so diam(GF(R))2\text{diam}(GF(R))\leqslant 2 by Theorem 6.7. Also note that from the above, if II and JJ do not have an edge between them, then we have the path I(0)JI\leftrightarrow(0)\leftrightarrow J.

Finally, for (3)(1)(3)\Longrightarrow(1), we suppose that IRI\subset R and consider the pair of ideals II and (0)(0). In either case there is an edge between them, and so II is finitely generated. ∎

We now give, in sequence, theorem for the behavior of GF()GF(-) for polynomial extensions, power series extensions, and homomorphic images. For the statement of these theorems, the case n=0n=0 corresponds to the case in which RR is a field (and the graph being a single vertex).

Theorem 6.15.

Suppose RR is a commutative ring with 1. If diam(GF(R))=\text{diam}(GF(R))=\infty then diam(GF(R[x1,x2,,xm]))=4\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{m}]))=4. If diam(GF(R))=n<\text{diam}(GF(R))=n<\infty then diam(GF(R))diam(GF(R[x1,x2,,xm]))\text{diam}(GF(R))\leqslant\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{m}])) and the following hold.

  1. (1)

    If n1n\leqslant 1, then diam(GF(R[x1,x2,,xm]))=2\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{m}]))=2.

  2. (2)

    If RR is Noetherian, then diam(GF(R[x1,x2,,xm]))=2\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{m}]))=2.

  3. (3)

    If n=4n=4, then diam(GF(R[x1,x2,,xm]))=4\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{m}]))=4.

Proof.

For the first statement, if diam(GF(R))=\text{diam}(GF(R))=\infty then RR is quasilocal with maximal ideal 𝔐\mathfrak{M} that is not finitely generated. Now note that R[x1,x2,,xn]/(x1,x2,,xn)RR[x_{1},x_{2},\cdots,x_{n}]/(x_{1},x_{2},\cdots,x_{n})\cong R. Since 𝔍(R)=𝔐\mathfrak{J}(R)=\mathfrak{M} is not contained in a finitely generated ideal, we have produced an ideal, namely (x1,x2,,xn)(x_{1},x_{2},\cdots,x_{n}) such that 𝔍(R[x1,x2,,xn]/(x1,x2,,xn))\mathfrak{J}(R[x_{1},x_{2},\cdots,x_{n}]/(x_{1},x_{2},\cdots,x_{n})) is not contained in a finitely generated ideal and so by Corollary 6.8, diam(GF(R[x1,x2,,xn]))=4\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{n}]))=4.

For the case diam(GF(R))=n<\text{diam}(GF(R))=n<\infty, we first note that in the case that diam(GF(R))2\text{diam}(GF(R))\leqslant 2 the fact that R[x1,x2,,xn]R[x_{1},x_{2},\cdots,x_{n}] is not chained shows by application of Theorem 6.3, that diam(GF(R[x1,x2,,xn]))2\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{n}]))\geqslant 2. If diam(GF(R))=3\text{diam}(GF(R))=3, then there is a maximal ideal 𝔐\mathfrak{M} that is not finitely generated. Since the ideal (𝔐,x1,x2,,xn)R[x1,x2,,xn](\mathfrak{M},x_{1},x_{2},\cdots,x_{n})\subset R[x_{1},x_{2},\cdots,x_{n}] is not finitely generated, diam(GF(R[x1,x2,,xn]))3\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{n}]))\geqslant 3. Finally, if diam(GF(R))=4\text{diam}(GF(R))=4, Corollary 6.8 gives that there is an ideal IRI\subset R such that 𝔍(R/I)\mathfrak{J}(R/I) is not contained in a finitely generated ideal. As R[x1,x2,,xn]/(I,x1,x2,,xn)R/IR[x_{1},x_{2},\cdots,x_{n}]/(I,x_{1},x_{2},\cdots,x_{n})\cong R/I, Corollary 6.8 again applies and diam(GF(R[x1,x2,,xn]))=4\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{n}]))=4.

For (1)(1), if n=1n=1 then by Theorem 6.3, RR is Noetherian and chained (this statement also holds in the case where n=0n=0 and RR is a field). Hence R[x1,x2,,xn]R[x_{1},x_{2},\cdots,x_{n}] is Noetherian but has infinitely many maximal ideals (so is not chained), therefore by Theorem 6.7 diam(GF(R[x1,x2,,xn]))=2\text{diam}(GF(R[x_{1},x_{2},\cdots,x_{n}]))=2.

For (2)(2), if RR is Noetherian, then R[x1,x2,,xn]R[x_{1},x_{2},\cdots,x_{n}] is Noetherian and the result follows as a porism of the proof of (1)(1).

(3)(3) is now immediate. ∎

Example 6.16.

Let R=𝔽[x,yx,yx2,]\displaystyle R=\mathbb{F}[x,\frac{y}{x},\frac{y}{x^{2}},\cdots] where 𝔽\mathbb{F} is a field and let V=R𝔐V=R_{\mathfrak{M}} where 𝔐=(x,yx,yx2,)\displaystyle\mathfrak{M}=(x,\frac{y}{x},\frac{y}{x^{2}},\cdots). VV is a 2-dimensional discrete valuation domain with principal maximal ideal (and hence has the property that all of its maximal ideals are finitely generated); its prime spectrum is 𝔐V=(x)𝔓=(y,yx,yx2,)(0)\mathfrak{M}V=(x)\supset\mathfrak{P}=(y,\frac{y}{x},\frac{y}{x^{2}},\cdots)\supset(0). But if we consider the polynomial ring V[t]V[t], it is easy to see that the ideal 𝔑:=(𝔓[t],xt+1)\mathfrak{N}:=(\mathfrak{P}[t],xt+1) is maximal, but not finitely generated. Note that the ideal (𝔓2[t],xt+1)(\mathfrak{P}^{2}[t],xt+1) has radical 𝔑\mathfrak{N} and hence V[t]/(𝔓2[t],xt+1)V[t]/(\mathfrak{P}^{2}[t],xt+1) has Jacobson radical 𝔑/(𝔓2[t],xt+1)\mathfrak{N}/(\mathfrak{P}^{2}[t],xt+1) but 𝔑\mathfrak{N} is not finitely generated over (𝔓2[t],xt+1)(\mathfrak{P}^{2}[t],xt+1). Thus diam(GF(V))=2\text{diam}(GF(V))=2 but diam(GF(V[t]))=4\text{diam}(GF(V[t]))=4. These details will also follow from the following.

From a more global perspective (in contrast with the class of almost Dedekind domains discussed in Proposition 6.13), the condition of Corollary 6.8 appears to make the situation where the diameter of GF(R)GF(R) is precisely 33 a reasonably rare occurrence. For example, we consider the class of SFT (for strong finite type) rings introduced by J. Arnold in [3] as a generalization of Noetherian rings useful in the study of the dimension of power series rings. An ideal IRI\subseteq R is said to be SFT if there is a finitely generated ideal BIB\subseteq I and a fixed natural number mm such that xmBx^{m}\in B for all xIx\in I; we say the ring, RR, is SFT if all of its ideals are SFT.

Proposition 6.17.

Let RR be SFT, or more generally, any ring with the property that each maximal ideal is the radical of a finitely generated ideal. Then diam(GF(R))3\text{diam}(GF(R))\neq 3.

Proof.

We can assume that there is a maximal ideal 𝔐\mathfrak{M} that is not finitely generated. By assumption, 𝔐\mathfrak{M} is the radical of a finitely generated ideal B𝔐B\subset\mathfrak{M}. Note that 𝔍(R/B)=𝔍(B)/B=𝔐/B\mathfrak{J}(R/B)=\mathfrak{J}(B)/B=\mathfrak{M}/B and is maximal. If 𝔐/B\mathfrak{M}/B is finitely generated, the fact that BB is finitely generated implies that 𝔐\mathfrak{M} is finitely generated. Hence 𝔍(R/B)\mathfrak{J}(R/B) is not contained in a finitely generated ideal and so by Corollary 6.8, diam(GF(R))=4\text{diam}(GF(R))=4. ∎

It is worth noting that diam(GF(R))=1,2,4,\text{diam}(GF(R))=1,2,4,\infty can occur for SFT rings. Also, Proposition 6.17 is another route to Example 6.16. Indeed, the domain VV from this example is a 22-dimensional SFT valuation domain and the results of [15] show that dim(V[t][[y]])\text{dim}(V[t][[y]]) is of finite Krull dimension and hence V[t]V[t] must be SFT. Since the maximal ideal 𝔑\mathfrak{N} is not finitely generated, diam(GF(V[t]))=4\text{diam}(GF(V[t]))=4.

Theorem 6.18.

If RR is a commutative ring with 1 and diam(GF(R))=n\text{diam}(GF(R))=n then the following hold.

  1. (0)

    If n=0n=0, then diam(GF(R[[x]]))=1\text{diam}(GF(R[[x]]))=1.

  2. (1)

    If n=1n=1, then diam(GF(R[[x1,x2,,xm]]))=2\text{diam}(GF(R[[x_{1},x_{2},\cdots,x_{m}]]))=2.

  3. (2)

    If n=2n=2, then diam(GF(R[[x1,x2,,xm]]))=2\text{diam}(GF(R[[x_{1},x_{2},\cdots,x_{m}]]))=2.

  4. (3)

    If n=3,4,n=3,4,\infty, then diam(GF(R[[x1,x2,,xm]]))=n\text{diam}(GF(R[[x_{1},x_{2},\cdots,x_{m}]]))=n.

Proof.

For (0)(0) we note that if RR is a field then R[[x]]R[[x]] is a Noetherian valuation domain.

For (1)(1), we have by Theorem 6.3 that RR is Noetherian and quasilocal and hence so is R[[x1,x2,,xn]]R[[x_{1},x_{2},\cdots,x_{n}]] (but not chained if RR is not a field) and so diam(GF(R[[x1,x2,,xn]]))=2\text{diam}(GF(R[[x_{1},x_{2},\cdots,x_{n}]]))=2 by Theorem 6.3.

(2)(2) is similar to the previous. Since each maximal ideal is finitely generated and the maximal ideals of R[[x1,x2,,xn]]R[[x_{1},x_{2},\cdots,x_{n}]] are of the form (𝔐,x1,x2,,xn)(\mathfrak{M},x_{1},x_{2},\cdots,x_{n}), we have that all maximal ideals of R[[x1,x2,,xn]]R[[x_{1},x_{2},\cdots,x_{n}]] are finitely generated as well.

For (3)(3), we first consider n=3n=3. Suppose that RR has the property that for all IRI\subset R, 𝔍(R/I)\mathfrak{J}(R/I) is contained in a finitely generated ideal. Now consider an ideal AR[[x]]A\subset R[[x]] and let A0={f(0)|f(x)A}A_{0}=\{f(0)|f(x)\in A\}. An easy computation shows that 𝔍(A)\mathfrak{J}(A) is precisely equal to (𝔍(A0),x)(\mathfrak{J}(A_{0}),x) where 𝔍(A0)\mathfrak{J}(A_{0}) is the Jacobson radical of A0A_{0} in RR. Note that there is an ideal FRF\subset R such that FF is finitely generated over 𝔍(A0)\mathfrak{J}(A_{0}) and A0A_{0} (that is, F/A0F/A_{0} is a finitely generated ideal containing 𝔍(A0)/A0=𝔍(R/A0)\mathfrak{J}(A_{0})/A_{0}=\mathfrak{J}(R/A_{0})). We conclude that (F,x)(F,x) is finitely generated over 𝔍(A)=(𝔍(A0),x)\mathfrak{J}(A)=(\mathfrak{J}(A_{0}),x) and with this in hand, we will show that (F,x)(F,x) is finitely generated over AA. To this end, we let (y1,y2,,yn)(y_{1},y_{2},\cdots,y_{n}) be the generators of FF over A0A_{0} and for each 1in1\leqslant i\leqslant n, find ai(x)Aa_{i}(x)\in A such that ai(0)=yia_{i}(0)=y_{i}. Now suppose that αp1(x)+xp2(x)(F,x)\alpha p_{1}(x)+xp_{2}(x)\in(F,x), with αF\alpha\in F and p1(x),p2(x)R[[x]]p_{1}(x),p_{2}(x)\in R[[x]]. Since xx is a generator, we can assume, without loss of generality, that p1:=p1(x)Rp_{1}:=p_{1}(x)\in R. Since αp1F\alpha p_{1}\in F, we can write αp1=r1y1+r2y2++rnyn+a0\alpha p_{1}=r_{1}y_{1}+r_{2}y_{2}+\cdots+r_{n}y_{n}+a_{0} with riR,a0A0r_{i}\in R,a_{0}\in A_{0}. Note that αp1+xp2(x)=r1a1(x)+r2a2(x)++rnan(x)+a(x)+xg(x)\alpha p_{1}+xp_{2}(x)=r_{1}a_{1}(x)+r_{2}a_{2}(x)+\cdots+r_{n}a_{n}(x)+a(x)+xg(x) where a(x)Aa(x)\in A with a(0)=a0a(0)=a_{0} and g(x)R[[x]]g(x)\in R[[x]]. So (F,x)(F,x) is generated over AA by the elements ai(x)a_{i}(x) and xx. Hence by Corollary 6.8 diam(R[[x]])=3\text{diam}(R[[x]])=3. The multivariable case follows immediately by induction.

If n=4n=4 then there is an ideal IRI\subset R such that 𝔍(R/I)\mathfrak{J}(R/I) is not contained in a finitely generated ideal. Note that as R[[x]]/(I,x)R/IR[[x]]/(I,x)\cong R/I this property persists in the power series extension. Again, induction completes this.

Finally, if n=n=\infty, RR is quasilocal with maximal ideal 𝔐\mathfrak{M} that is not finitely generated by Theorem 6.3. Since R[[x1,x2,,xn]]R[[x_{1},x_{2},\cdots,x_{n}]] is quasilocal with maximal ideal (𝔐,x1,x2,,xn)(\mathfrak{M},x_{1},x_{2},\cdots,x_{n}) the diameter of GF(R[[x1,x2,,xn]])GF(R[[x_{1},x_{2},\cdots,x_{n}]]) is also infinite. ∎

We round this out by recording behavior in homomorphic images.

Theorem 6.19.

Let RR be a commutative ring with 1 and IRI\subset R an ideal. If diam(GF(R))=n\text{diam}(GF(R))=n then the following hold.

  1. (1)

    If n2n\leqslant 2, then diam(GF(R/I))n\text{diam}(GF(R/I))\leqslant n.

  2. (2)

    If n=3n=3, then diam(GF(R/I))n\text{diam}(GF(R/I))\leqslant n.

  3. (3)

    If n=4n=4, then diam(GF(R/I))n\text{diam}(GF(R/I))\leqslant n or diam(GF(R/I))=\text{diam}(GF(R/I))=\infty.

Proof.

If n2n\leqslant 2 it is necessarily true that the maximal ideals of RR are finitely generated and hence the same is true of R/IR/I.

For the second statement, we suppose that diam(GF(R))=3\text{diam}(GF(R))=3, and so in particular, for all IRI\subset R, 𝔍(R/I)\mathfrak{J}(R/I) is contained in a finitely generated ideal. So if J/IJ/I is an ideal of R/IR/I then (R/I)/(J/I)R/J(R/I)/(J/I)\cong R/J and so GF(R/I)GF(R/I) has diameter of no more than 33.

For (3)(3), if diam(GF(R/I))\text{diam}(GF(R/I)) is finite, then the first statement is clear. Note however, that if RR is a ring with a maximal ideal 𝔐\mathfrak{M} that is not finitely generated, and I𝔐I\subset\mathfrak{M} has the property that I=𝔐\sqrt{I}=\mathfrak{M} and 𝔐/I\mathfrak{M}/I is not finitely generated, then R/IR/I is quasilocal and hence its graph is not connected by Theorem 6.3. For a concrete example of this, let VV be a 11-dimensional nondiscrete valuation domain. By Theorem 6.3 diam(GF(V))=\text{diam}(GF(V))=\infty and diam(GF(V[x]))=4\text{diam}(GF(V[x]))=4 by Theorem 6.15; since VV[x]/(x)V\cong V[x]/(x) we have our example. ∎

Remark 6.20.

As a final remark, we point out that it would be interesting to know if the property that diam(GF(R))=3\text{diam}(GF(R))=3 is stable under polynomial extensions. More generally, a more complete understanding of stability properties of the diameter of the graphs GF()GF(-) under polynomial extensions for n=2,3n=2,3 would be desirable.

References

  • [1] David F. Anderson and John D. LaGrange. Some remarks on the compressed zero-divisor graph. J. Algebra, 447:297–321, 2016.
  • [2] David F. Anderson and Philip S. Livingston. The zero-divisor graph of a commutative ring. J. Algebra, 217(2):434–447, 1999.
  • [3] J. T. Arnold. Krull dimension in power series rings. Trans. Amer. Math. Soc., 177:299–304, 1973.
  • [4] M. Axtell, N. Baeth, and J. Stickles. Cut structures in zero-divisor graphs of commutative rings. J. Commut. Algebra, 8(2):143–171, 2016.
  • [5] Michael Axtell, James Coykendall, and Joe Stickles. Zero-divisor graphs of polynomials and power series over commutative rings. Comm. Algebra, 33(6):2043–2050, 2005.
  • [6] István Beck. Coloring of commutative rings. J. Algebra, 116(1):208–226, 1988.
  • [7] M. Behboodi and Z. Rakeei. The annihilating-ideal graph of commutative rings I. J. Algebra Appl., 10(4):727–739, 2011.
  • [8] Jason Greene Boynton and Jim Coykendall. On the graph of divisibility of an integral domain. Canad. Math. Bull., 58(3):449–458, 2015.
  • [9] Jim Coykendall and Jack Maney. Irreducible divisor graphs. Comm. Algebra, 35(3):885–895, 2007.
  • [10] Frank DeMeyer and Lisa DeMeyer. Zero divisor graphs of semigroups. J. Algebra, 283(1):190–198, 2005.
  • [11] Robert Gilmer and Raymond C. Heitmann. On Pic(R[X]){\rm Pic}(R[X]) for RR seminormal. J. Pure Appl. Algebra, 16(3):251–257, 1980.
  • [12] Robert W Gilmer. Multiplicative ideal theory. Number 12. Queen’s University Kinston, Ontario, 1968.
  • [13] Jürgen Herzog, Somayeh Moradi, and Masoomeh Rahimbeigi. The edge ideal of a graph and its splitting graphs. J. Commut. Algebra, 14(1):27–35, 2022.
  • [14] Takayuki Hibi, Kyouko Kimura, Kazunori Matsuda, and Akiyoshi Tsuchiya. Regularity and aa-invariant of Cameron-Walker graphs. J. Algebra, 584:215–242, 2021.
  • [15] B. G. Kang and M. H. Park. Krull dimension of mixed extensions. J. Pure Appl. Algebra, 213(10):1911–1915, 2009.
  • [16] K. Alan Loper. Almost Dedekind domains which are not Dedekind. In Multiplicative ideal theory in commutative algebra, pages 279–292. Springer, New York, 2006.
  • [17] Thomas G. Lucas. The clique ideal property. J. Commut. Algebra, 10(4):499–546, 2018.
  • [18] Aron Simis, Wolmer V. Vasconcelos, and Rafael H. Villarreal. On the ideal theory of graphs. J. Algebra, 167(2):389–416, 1994.
  • [19] Richard G. Swan. On seminormality. J. Algebra, 67(1):210–229, 1980.
  • [20] Rafael H. Villarreal. Cohen-Macaulay graphs. Manuscripta Math., 66(3):277–293, 1990.
  • [21] Rafael H. Villarreal. Rees algebras of edge ideals. Comm. Algebra, 23(9):3513–3524, 1995.