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

Properties of the Toric Rings of a Chordal Bipartite Family of Graphs

Laura Ballard Mathematics Department, Syracuse University, Syracuse, NY 13244, U.S.A. [email protected]
Abstract.

This work concerns the study of properties of a group of Koszul algebras coming from the toric ideals of a chordal bipartite infinite family of graphs (alternately, these rings may be interpreted as coming from determinants of certain ladder-like structures). We determine a linear system of parameters for each ring and explicitly determine the Hilbert series for the resulting Artinian reduction. As corollaries, we obtain the multiplicity and regularity of the original rings. This work extends results easily derived from lattice theory for a subfamily coming from a two-sided ladder to a family where, as we show, lattice theory no longer applies in any obvious way and includes constructive proofs which may be useful in future study of these rings and others.

Key words and phrases:
Koszul algebra, toric ring, graph, chordal bipartite, Castelnuovo-Mumford regularity, multiplicity
2020 Mathematics Subject Classification:
05E40, 13C15, 13D02, 13F65, 13H15, 16S37
L. Ballard was partially supported by the National Science Foundation (DMS-1003384).

1. Introduction

In recent decades, there has been a growing interest in the investigation of algebraic invariants associated to combinatorial structures. Toric ideals of graphs (and the associated edge rings), a special case of the classical notion of a toric ideal, have been studied by various authors with regard to invariants such as depth, dimension, projective dimension, regularity, graded Betti numbers, Hilbert series, and multiplicity, usually for particular families of graphs (see for example [2, 3, 5, 7, 8, 9, 10, 12, 14, 16, 17, 18, 19, 22, 23, 26]). We note in Remarks 2.6 and 2.14 that the family we consider does not overlap at all or for large nn with those considered in [5], [8], [9], and [23]; it is more obviously distinct from other families that have been studied. We think it fitting to mention that the recent book by Herzog, Hibi, and Ohsugi ([15]) also investigates toric ideals of graphs as well as binomial ideals coming from other combinatorial structures.

In this work, we consider a family of graphs with iterated subfamilies and develop algebraic properties of the toric rings associated to the family which depend only on the number of vertices (equivalently, the number of edges) in the associated graphs. In the development of this project, we were particularly inspired by the work of Jennifer Biermann, Augustine O’Keefe, and Adam Van Tuyl in [3], where they establish a lower bound for the regularity of the toric ideal of any finite simple graph and an upper bound for the regularity of the toric ideal of a chordal bipartite graph. Our goal is to construct as “simple” a family of graphs as possible that still yields interesting toric ideals. It is our hope that our process and results will lead to further generalizations of properties of toric ideals for other (perhaps broader) families of graphs, or for graphs containing or arising from such graphs.

Herein, we introduce the infinite family \mathcal{F} of chordal bipartite graphs GntG_{n}^{t}, where nn determines the number of edges and vertices and tt determines the structure of the graph, and establish some algebraic properties of the toric rings R(n,t)R(n,t) associated to the graphs GntG_{n}^{t}. The use of bipartite graphs makes each R(n,t)R(n,t) normal and Cohen-Macaulay by [25] and [15]; we use the latter in Section 3. Our main results prove to be independent of tt and depend only on nn.

In Section 2, we construct the family \mathcal{F} of graphs GntG_{n}^{t} from a family of ladder-like structures LntL_{n}^{t} so that the toric ideals of the GntG_{n}^{t} are generalized determinantal ideals of the LntL_{n}^{t}. The ladder-like structures associated to a subfamily 1\mathcal{F}_{1}\subset\mathcal{F}, introduced in Example 2.4, are in fact two-sided ladders (for large nn), so that the family of rings R(n,t)R(n,t) is a generalization of the family of ladder determinantal rings coming from 1\mathcal{F}_{1}. While the rings arising from 1\mathcal{F}_{1} come from a distributive lattice and have easily derived properties (see for example [15]), we show that the rings associated to \mathcal{F} do not naturally arise from any lattice in general, and merit closer study.

In Section 3, we establish some algebraic properties of the R(n,t)R(n,t), particularly Krull dimension, projective dimension, multiplicity, and regularity. To do so, we prove that the determinantal generators of the defining ideal IGntI_{G_{n}^{t}} are a Gröbner basis (it follows immediately from [15] that R(n,t)R(n,t) is Koszul) and work with the initial ideal in>IGnt\text{in}_{>}I_{G_{n}^{t}}. We also develop a system of parameters Xn¯\overline{X_{n}} that allows us to work with Artinian reductions in part of our treatment, and their Hilbert series.

Our first result establishes the Krull dimension of the toric ring R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}}, where the ring S(n)=k[x0,x2,x3,,x2n+3,x2n+4]S(n)=k[x_{0},x_{2},x_{3},\ldots,x_{2n+3},x_{2n+4}] is the polynomial ring over the edges of GntG_{n}^{t} and IGntI_{G_{n}^{t}} is the toric ideal of GntG_{n}^{t}.

Theorem 1.1 (Theorem 3.4).

The dimension of R(n,t)R(n,t) is

dimR(n,t)=n+3.\dim R(n,t)=n+3.

As a corollary, since R(n,t)R(n,t) comes from a bipartite graph and is hence Cohen-Macaulay (Corollary 2.16), we obtain the projective dimension of R(n,t)R(n,t).

Corollary 1.2 (Corollary 3.5).

The projective dimension of R(n,t)R(n,t) over S(n)S(n) is

pdS(n)R(n,t)=n+1.\textnormal{pd}_{S(n)}R(n,t)=n+1.

We then develop a linear system of parameters for R(n,t)R(n,t), using differences of elements on antidiagonals of the ladder-like structure LntL_{n}^{t}.

Proposition 1.3 (Proposition 3.10).

Let R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}}. Then the image of

Xn=x0,x2x3,x4x5,,x2nx2n+1,x2n+2x2n+3,x2n+4X_{n}=x_{0},x_{2}-x_{3},x_{4}-x_{5},\ldots,x_{2n}-x_{2n+1},x_{2n+2}-x_{2n+3},x_{2n+4}

in R(n,t)R(n,t) is a system of parameters for R(n,t)R(n,t).

Since R(n,t)R(n,t) is Cohen-Macaulay, the linear system of parameters above is a regular sequence (Corollary 3.12).

With the aim of obtaining the multiplicity and regularity of R(n,t)R(n,t), we form an Artinian quotient of R(n,t)R(n,t) by the regular sequence above and call it R(n,t)^\widehat{R(n,t)}. We note that R(n,t)^\widehat{R(n,t)} does not denote the completion, and explain the choice of notation in Definition 3.7.

Using a convenient vector space basis for R(n,t)^\widehat{R(n,t)} established in Lemma 3.13, we show the coefficients of the Hilbert series for R(n,t)^\widehat{R(n,t)}.

Theorem 1.4 (Theorem 3.16).

If R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}} and R(n,t)^R(n,t)/(Xn¯)\widehat{R(n,t)}\cong R(n,t)/(\overline{X_{n}}), we have

dimk(R(n,t)^)i={1i=01i!j=1i(n+j2(i1))i1.{\displaystyle\dim_{k}(\widehat{R(n,t)})_{i}=\begin{cases}1&i=0\\ {\displaystyle\frac{1}{i!}\prod_{j=1}^{i}(n+j-2(i-1))}&i\geq 1.\end{cases}}

In particular, dimk(R(n,t)^)i=0\dim_{k}(\widehat{R(n,t)})_{i}=0 when i>n/2+1i>n/2+1.

As a corollary, we obtain the regularity of R(n,t)R(n,t), which is equal to the top nonzero degree of R(n,t)^\widehat{R(n,t)}.

Corollary 1.5 (Corollary 3.18).

For GntG_{n}^{t}\in\mathcal{F},

regR(n,t)=n/2+1.\operatorname{reg}R(n,t)=\left\lfloor n/2\right\rfloor+1.

We include an alternate graph-theoretic proof of the result above at the end of this work. Beginning with an upper bound from [3] (or equivalently for our purposes, one from [14]) and then identifying the initial ideal in>IGnt\text{in}_{>}I_{G_{n}^{t}} with the edge ideal of a graph, we use results from [4] (allowing us to use in>IGnt\text{in}_{>}I_{G_{n}^{t}} instead of IGntI_{G_{n}^{t}}) and then [13] for a lower bound which agrees with our upper bound.

From a recursion established in Lemma 3.15, we go on to prove a Fibonacci relationship between the lengths of the Artinian rings R(n,t)^\widehat{R(n,t)} in Proposition 3.19, and obtain the multiplicity of R(n,t)R(n,t) as a corollary. In the following, we drop tt for convenience.

Corollary 1.6 (Corollary 3.21).

For n2n\geq 2, there is an equality of multiplicities

e(R(n))=e(R(n1))+e(R(n2)).e(R(n))=e(R(n-1))+e(R(n-2)).

In particular,

e(R(n))=F(n+3)=(1+5)n+3(15)n+32n+35.e(R(n))=F\left(n+3\right)=\frac{(1+\sqrt{5})^{n+3}-(1-\sqrt{5})^{n+3}}{2^{n+3}\sqrt{5}}.

For more background, detail, and motivation, we refer the reader to [1], but note that different notation and indexing conventions have been employed in this work. Throughout, kk is a field.

Acknowledgements. Macaulay2 [11] was used for computation and hypothesis formation. We would like to thank Syracuse University for its support and hospitality and Claudia Miller for her valuable input on the original project in [1] and this condensed version. We also acknowledge the partial support of an NSF grant.

2. The Family of Toric Rings

In the following, we define a family of toric rings R(n,t)R(n,t) coming from an iterative chordal bipartite family of graphs, \mathcal{F}. We show that although one subfamily of these rings comes from join-meet ideals of a (distributive) lattice and has some easily derived algebraic invariants, this is not true in general. The reader may find the definition of the toric ideal of a graph in Section 2.2, when it becomes relevant to the discussion. We recall for the reader that a chordal bipartite graph is a bipartite graph in which every cycle of length greater than or equal to six has a chord.

2.1. The Family \mathcal{F} of Graphs

Below, we define the family \mathcal{F} of chordal bipartite graphs iteratively from a family of ladder-like structures LntL_{n}^{t}. We note that the quantities involved in the following definition follow patterns as follows:

nn n/2+2\left\lfloor{n/2}\right\rfloor+2 n/2+2\left\lceil{n/2}\right\rceil+2
0 2 2
1 2 3
2 3 3
3 3 4
Definition 2.1.

For each n0n\geq 0 and each t𝔽2n+1t\in\mathbb{F}_{2}^{n+1}, we construct a ladder-like structure LntL_{n}^{t} with (n/2+2)(\left\lfloor{n/2}\right\rfloor+2) rows and (n/2+2)(\left\lceil{n/2}\right\rceil+2) columns and with nonzero entries in the set {x0,x2,x3,,x2n+4}\{x_{0},x_{2},x_{3},\ldots,x_{2n+4}\}. To do so, we use the notation t^𝔽2n\widehat{t}\in\mathbb{F}_{2}^{n} for the first nn entries of tt, that is, all except the last entry. The construction is as follows, where throughout, indices of entries in LntL_{n}^{t} are strictly increasing from left to right in each row and from top to bottom in each column. We note that LntL_{n}^{t} does not depend on tt for n<2n<2, but does for n2n\geq 2.

  • For n=0n=0, the ladder-like structure L00=L01L_{0}^{0}=L_{0}^{1} is

    x0x2x3x4\begin{matrix}x_{0}&x_{2}\\ x_{3}&x_{4}\end{matrix}
  • For n=1n=1, to create L1tL_{1}^{t} (regardless of what tt is in 𝔽22\mathbb{F}_{2}^{2}), we add another column with the entries x5x_{5} and x6x_{6} to the right of L0t^L_{0}^{\widehat{t}} to obtain

    x0x2x5x3x4x6\begin{matrix}x_{0}&x_{2}&x_{5}\\ x_{3}&x_{4}&x_{6}\end{matrix}
  • For 2n0mod2(1mod2)2\leq n\equiv 0\mod 2(\equiv 1\mod 2), to create LntL_{n}^{t}, we add another row (column) with the entries x2n+3,x2n+4x_{2n+3},x_{2n+4} below (to the right of) Ln1t^L_{n-1}^{\widehat{t}} in the following way:

    • \circ

      The entry x2n+4x_{2n+4} is in the ultimate row and column, row n/2+2\left\lfloor{n/2}\right\rfloor+2 and column n/2+2\left\lceil{n/2}\right\rceil+2.

    • \circ

      The entry x2n+3x_{2n+3} is in the new row (column) in a position directly below (to the right of) another nonzero entry in LntL_{n}^{t}.

      • *

        If the last entry of tt is 0, x2n+3x_{2n+3} is directly beneath (to the right of) the first nonzero entry in the previous row (column).

      • *

        If the last entry of tt is 11, x2n+3x_{2n+3} is directly beneath (to the right of) the second nonzero entry in the previous row (column).

In this way, the entries in tt determine the choice at each stage for the placement of x2n+3x_{2n+3}.

Remark 2.2.

We note a few things about this construction for n0mod2n\equiv 0\mod 2 (1mod2\equiv 1\mod 2), which may be examined in the examples below:

  • We note that x2n+4x_{2n+4} is directly beneath (to the right of) x2n+2x_{2n+2}.

  • We note that the only entries in row n/2+1\left\lfloor{n/2}\right\rfloor+1 (column n/2+1\left\lceil{n/2}\right\rceil+1) of Ln1t^L_{n-1}^{\widehat{t}} are x2n1x_{2n-1}, x2nx_{2n}, and x2n+2x_{2n+2}, so that the choices listed for placement of x2n+3x_{2n+3} are the only cases. In particular, tn+1=0t_{n+1}=0 if and only if x2n+3x_{2n+3} is directly beneath (to the right of) x2n1x_{2n-1}, and tn+1=1t_{n+1}=1 if and only if x2n+3x_{2n+3} is directly beneath (to the right of) x2nx_{2n}.

  • Finally, we note that the only entries in column n/2+2\left\lceil{n/2}\right\rceil+2 (row n/2+2\left\lfloor{n/2}\right\rfloor+2) of LntL_{n}^{t} are x2n+1x_{2n+1}, x2n+2x_{2n+2}, and x2n+4x_{2n+4}, and that the only entries in row n/2+2\left\lfloor{n/2}\right\rfloor+2 (column n/2+2\left\lceil{n/2}\right\rceil+2) of LntL_{n}^{t} are x2n+3x_{2n+3} and x2n+4x_{2n+4}.

Example 2.3.

We have

L2(1,1,1)=x0x2x5x3x4x6x7x8L2(0,0,0)=x0x2x5x3x4x6x7x8L_{2}^{(1,1,1)}=\begin{matrix}x_{0}&x_{2}&x_{5}\\ x_{3}&x_{4}&x_{6}\\ &x_{7}&x_{8}\end{matrix}\hskip 56.9055ptL_{2}^{(0,0,0)}=\begin{matrix}x_{0}&x_{2}&x_{5}\\ x_{3}&x_{4}&x_{6}\\ x_{7}&&x_{8}\end{matrix}

In either of the cases above, we could go on to construct L3t^L_{3}^{\widehat{t}} and L4tL_{4}^{t} in the following way: For n=3n=3, place x10x_{10} to the right of x8x_{8} and place x9x_{9} to the right of either x5x_{5} or x6x_{6}, depending whether the last entry of t^\widehat{t} is 0 or 11, respectively. Then for n=4n=4, place x12x_{12} below x10x_{10} and place x11x_{11} below either x7x_{7} or x8x_{8}, depending whether the last entry of tt is 0 or 11, respectively.

Example 2.4.

In fact, when the entries of tt are all ones, we see that Ln(1,1,,1)L_{n}^{(1,1,\ldots,1)} has a ladder shape (is a two-sided ladder for n3)n\geq 3), shown below in the case when 2n0mod22\leq n\equiv 0\mod 2:

x0x2x5x3x4x6x9x7x8x10x13x11x12x14x17x15x16x18x21x19x20x22x25x23x24x26x27x28x2n+1x2n+2x2n+3x2n+4.\begin{matrix}x_{0}&x_{2}&x_{5}&\\ x_{3}&x_{4}&x_{6}&x_{9}&\\ &x_{7}&x_{8}&x_{10}&x_{13}&\\ &&x_{11}&x_{12}&x_{14}&x_{17}&\\ &&&x_{15}&x_{16}&x_{18}&x_{21}&\\ &&&&x_{19}&x_{20}&x_{22}&x_{25}&\\ &&&&&x_{23}&x_{24}&x_{26}&\ddots&\\ &&&&&&x_{27}&x_{28}&\ddots&x_{2n+1}\\ &&&&&&&\ddots&\ddots&x_{2n+2}\\ &&&&&&&&x_{2n+3}&x_{2n+4}.\end{matrix}

We denote the subfamily of graphs coming from t=(1,1,,1)t=(1,1,\ldots,1) by 1\mathcal{F}_{1}\subset\mathcal{F}.

When the entries of tt are all zeros, Ln(0,0,,0)L_{n}^{(0,0,\ldots,0)} has the following structure, shown below in the case when 2n0mod22\leq n\equiv 0\mod 2:

x0x2x5x9x13x17x21x25x2n+1x3x4x6x7x8x10x11x12x14x15x16x18x19x20x22x23x24x26x27x28x2n+2x2n+3x2n+4.\begin{matrix}x_{0}&x_{2}&x_{5}&x_{9}&x_{13}&x_{17}&x_{21}&x_{25}&\cdots&x_{2n+1}\\ x_{3}&x_{4}&x_{6}\\ x_{7}&&x_{8}&x_{10}\\ x_{11}&&&x_{12}&x_{14}\\ x_{15}&&&&x_{16}&x_{18}\\ x_{19}&&&&&x_{20}&x_{22}\\ x_{23}&&&&&&x_{24}&x_{26}\\ x_{27}&&&&&&&x_{28}&\ddots\\ \vdots&&&&&&&&\ddots&x_{2n+2}\\ x_{2n+3}&&&&&&&&&x_{2n+4}.\end{matrix}

For a more varied example, we have L16(1,0,1,0,1,1,0,0,1,1,1,0,0,0,1,0,0)L_{16}^{(1,0,1,0,1,1,0,0,1,1,1,0,0,0,1,0,0)} below:

x0x2x5x9x3x4x6x7x8x10x13x17x11x12x14x15x16x18x21x25x29x33x19x20x22x23x24x26x27x28x30x31x32x34x35x36.\begin{matrix}x_{0}&x_{2}&x_{5}&x_{9}&&&&&&\\ x_{3}&x_{4}&x_{6}&&&&&&&\\ &x_{7}&x_{8}&x_{10}&x_{13}&x_{17}&&&&\\ &&x_{11}&x_{12}&x_{14}&&&&&\\ &&x_{15}&&x_{16}&x_{18}&x_{21}&x_{25}&x_{29}&x_{33}\\ &&&&x_{19}&x_{20}&x_{22}&&&\\ &&&&&x_{23}&x_{24}&x_{26}&&\\ &&&&&x_{27}&&x_{28}&x_{30}&\\ &&&&&&&x_{31}&x_{32}&x_{34}\\ &&&&&&&x_{35}&&x_{36}.\end{matrix}
Definition 2.5.

If we associate a vertex to each row and each column and an edge to each nonzero entry of LntL_{n}^{t}, we have a finite simple connected bipartite graph GntG_{n}^{t}. The set VrV_{r} of vertices corresponding to rows and the set VcV_{c} of vertices corresponding to columns form a bipartition of the vertices of GntG_{n}^{t}. We say a graph GG is in \mathcal{F} if G=GntG=G_{n}^{t} for some n0n\geq 0 and some t𝔽2n+1t\in\mathbb{F}_{2}^{n+1}.

Remark 2.6.

We note that by construction GntG_{n}^{t} has no vertices of degree one, since each row and each column of LntL_{n}^{t} has more than one nonzero entry. This ensures that for large nn our family is distinct from that studied in [5], since a Ferrers graph with bipartitation V1V_{1} and V2V_{2} with no vertices of degree one must have at least two vertices in V1V_{1} of degree |V2||V_{2}| and at least two vertices in V2V_{2} of degree |V1||V_{1}|, impossible for our graphs when n3n\geq 3, as the reader may verify. We also use the fact that GntG_{n}^{t} has no vertices of degree one for an alternate proof of the regularity of R(n,t)R(n,t) at the end of this work.

Example 2.7.

When n=5n=5, G5(1,1,,1)1G_{5}^{(1,1,\ldots,1)}\in\mathcal{F}_{1} is

r1r1c1c1c2c2r2r2c3c3r3r3c4c4r4r4c5c5x0x_{0}x2x_{2}x5x_{5}x3x_{3}x4x_{4}x6x_{6}x9x_{9}x7x_{7}x8x_{8}x10x_{10}x13x_{13}x11x_{11}x12x_{12}x14x_{14}

We develop properties of the LntL_{n}^{t} which allow us to show in Section 2.2 that certain minors of the LntL_{n}^{t} are generators for the toric rings of the corresponding graphs GntG_{n}^{t}.

Definition 2.8.

For this work, a distinguished minor of LntL_{n}^{t} is a 22-minor involving only (nonzero) entries of the ladder-like structure LntL_{n}^{t}, coming from a 2×22\times 2 subarray of LntL_{n}^{t}.

Proposition 2.9.

For each i1i\geq 1 and each f𝔽2i+1f\in\mathbb{F}_{2}^{i+1}, the entry x2i+3x_{2i+3} and the entry x2i+4x_{2i+4} each appear in exactly two distinguished minors of LifL_{i}^{f}. For i0mod2(1mod2)i\equiv 0\mod 2(\equiv 1\mod 2), these minors are of the form

s2i:=x2i+1x2i+3xj2ix2i+4s_{2i}:=x_{2i+1}x_{2i+3}-x_{j_{2i}}x_{2i+4}

coming from the subarray

[xj2ix2i+1x2i+3x2i+4]([xj2ix2i+3x2i+1x2i+4])\begin{bmatrix}x_{j_{2i}}&x_{2i+1}\\ x_{2i+3}&x_{2i+4}\end{bmatrix}\hskip 56.9055pt\left(\begin{bmatrix}x_{j_{2i}}&x_{2i+3}\\ x_{2i+1}&x_{2i+4}\end{bmatrix}\right)

for some j2i{0,2,3,,2i2}j_{2i}\in\{0,2,3,\ldots,2i-2\} and

s2i+1:=x2i+2x2i+3xj2i+1x2i+4s_{2i+1}:=x_{2i+2}x_{2i+3}-x_{j_{2i+1}}x_{2i+4}

coming from the subarray

[xj2i+1x2i+2x2i+3x2i+4]([xj2i+1x2i+3x2i+2x2i+4])\begin{bmatrix}x_{j_{2i+1}}&x_{2i+2}\\ x_{2i+3}&x_{2i+4}\end{bmatrix}\hskip 56.9055pt\left(\begin{bmatrix}x_{j_{2i+1}}&x_{2i+3}\\ x_{2i+2}&x_{2i+4}\end{bmatrix}\right)

for some j2i+1{2i1,2i}j_{2i+1}\in\{2i-1,2i\}, and the only distinguished minor of LntL_{n}^{t} with indices all less than 55 is s1:=x2x3x0x4s_{1}:=x_{2}x_{3}-x_{0}x_{4}.

Proof.

The last statement is clear by Definition 2.1; we prove the remaining statements by induction on ii. For i=1i=1, we have the distinguished minors s2=x3x5x0x6s_{2}=x_{3}x_{5}-x_{0}x_{6} and s3=x4x5x2x6s_{3}=x_{4}x_{5}-x_{2}x_{6} coming from the subarrays

[x0x5x3x6]\begin{bmatrix}x_{0}&x_{5}\\ x_{3}&x_{6}\end{bmatrix}

and

[x2x5x4x6]\begin{bmatrix}x_{2}&x_{5}\\ x_{4}&x_{6}\end{bmatrix}

where j2=0{0}j_{2}=0\in\{0\} and j3=2{1,2}j_{3}=2\in\{1,2\}, so we have our base case. Now suppose the statement is true for ii with 1i<n1\leq i<n, and let n0mod2(1mod2)n\equiv 0\mod 2(\equiv 1\mod 2) and t𝔽2n+1t\in\mathbb{F}_{2}^{n+1}.

Case 1: If tn+1=0t_{n+1}=0, then by Remark 2.2, x2n+3x_{2n+3} is in the same column (row) as x2n1x_{2n-1}. By induction, we have the distinguished minor s2n2=x2n1x2n+1xj2n2x2n+2s_{2n-2}=x_{2n-1}x_{2n+1}-x_{j_{2n-2}}x_{2n+2} coming from the subarray

[xj2n2x2n+1x2n1x2n+2]([xj2n2x2n1x2n+1x2n+2]).\begin{bmatrix}x_{j_{2n-2}}&x_{2n+1}\\ x_{2n-1}&x_{2n+2}\end{bmatrix}\hskip 56.9055pt\left(\begin{bmatrix}x_{j_{2n-2}}&x_{2n-1}\\ x_{2n+1}&x_{2n+2}\end{bmatrix}\right).

Then in fact we have a subarray of the form

[xj2n2x2n+1x2n1x2n+2x2n+3x2n+4]([xj2n2x2n1x2n+3x2n+1x2n+2x2n+4]),\begin{bmatrix}x_{j_{2n-2}}&x_{2n+1}\\ x_{2n-1}&x_{2n+2}\\ x_{2n+3}&x_{2n+4}\end{bmatrix}\hskip 56.9055pt\left(\begin{bmatrix}x_{j_{2n-2}}&x_{2n-1}&x_{2n+3}\\ x_{2n+1}&x_{2n+2}&x_{2n+4}\end{bmatrix}\right),

so that we have the distinguished minors

s2n\displaystyle s_{2n} =\displaystyle= x2n+1x2n+3xj2n2x2n+4\displaystyle x_{2n+1}x_{2n+3}-x_{j_{2n-2}}x_{2n+4}
s2n+1\displaystyle s_{2n+1} =\displaystyle= x2n+2x2n+3x2n1x2n+4\displaystyle x_{2n+2}x_{2n+3}-x_{2n-1}x_{2n+4}

with

j2n=j2n2{0,2,3,,2n4}{0,2,3,,2n2}j_{2n}=j_{2n-2}\in\{0,2,3,\ldots,2n-4\}\subset\{0,2,3,\ldots,2n-2\}

by induction and with

j2n+1=2n1{2n1,2n}.j_{2n+1}=2n-1\in\{2n-1,2n\}.

Since the only entries in row n/2+2\left\lfloor{n/2}\right\rfloor+2 (column n/2+2\left\lceil{n/2}\right\rceil+2) of LntL_{n}^{t} are x2n+3x_{2n+3} and x2n+4x_{2n+4} and since the only entries in column n/2+2\left\lceil{n/2}\right\rceil+2 (row n/2+2\left\lfloor{n/2}\right\rfloor+2) of LntL_{n}^{t} are x2n+1x_{2n+1}, x2n+2x_{2n+2}, and x2n+4x_{2n+4} by Remark 2.2, these are the only distinguished minors of LntL_{n}^{t} containing either x2n+3x_{2n+3} or x2n+4x_{2n+4}, as desired.

Case 2 for tn+1=1t_{n+1}=1 is analogous and yields

j2n=j2n1{2n3,2n2}{0,2,3,,2n2}j_{2n}=j_{2n-1}\in\{2n-3,2n-2\}\subset\{0,2,3,\ldots,2n-2\}

and

j2n+1=2n{2n1,2n}.j_{2n+1}=2n\in\{2n-1,2n\}.

Definition 2.10.

Define the integers j2ij_{2i}, j2i+1j_{2i+1} for j2,,j2n+1j_{2},\ldots,j_{2n+1} as in the statement of Proposition 2.9. We note in the remark below some properties of the jkj_{k}.

Remark 2.11.

From the proof of Proposition 2.9, we note that j2=0j_{2}=0, j3=2j_{3}=2, and that for i2i\geq 2, we have the following:

ti+1\displaystyle t_{i+1} =\displaystyle= 0j2i=j2i2j2i+1=2i1\displaystyle 0\iff j_{2i}=j_{2i-2}\iff j_{2i+1}=2i-1
ti+1\displaystyle t_{i+1} =\displaystyle= 1j2i=j2i1j2i+1=2i.\displaystyle 1\iff j_{2i}=j_{2i-1}\iff j_{2i+1}=2i.

For the sake of later proofs, we extend the notion of jkj_{k} naturally to s1=x2x3x0x4s_{1}=x_{2}x_{3}-x_{0}x_{4} and say that j1=0j_{1}=0, and note the following properties of the jkj_{k} for 1k2n+11\leq k\leq 2n+1:

  • We have j2i{j2i2,j2i1}j_{2i}\in\{j_{2i-2},j_{2i-1}\} and j2i2i2j_{2i}\leq 2i-2. Indeed, for i=1i=1, j2=j1=0j_{2}=j_{1}=0, and for i2i\geq 2, this is clear from the statement above.

  • We have j2i+1{2i1,2i}j_{2i+1}\in\{2i-1,2i\}. Indeed, for i=0i=0, j1=0{1,0}j_{1}=0\in\{-1,0\}, for i=1i=1, j3=2{1,2}j_{3}=2\in\{1,2\}, and for i2i\geq 2, this follows from the statement above.

  • The j2ij_{2i} form a non-decreasing sequence. Indeed, for i2i\geq 2, either j2i=j2i2j_{2i}=j_{2i-2} or j2i=j2i12i3>2i4j2i2j_{2i}=j_{2i-1}\geq 2i-3>2i-4\geq j_{2i-2}.

Remark 2.12.

We also note from the proof above that the following is a subarray of LntL_{n}^{t} for all i0mod2(1mod2)i\equiv 0\mod 2(\equiv 1\mod 2) such that 1in1\leq i\leq n, which we use in the proof of the proposition below:

[xj2ix2i+1xj2i+1x2i+2x2i+3x2i+4]([xj2ixj2i+1x2i+3x2i+1x2i+2x2i+4])\begin{bmatrix}x_{j_{2i}}&x_{2i+1}\\ x_{j_{2i+1}}&x_{2i+2}\\ x_{2i+3}&x_{2i+4}\end{bmatrix}\hskip 56.9055pt\left(\begin{bmatrix}x_{j_{2i}}&x_{j_{2i+1}}&x_{2i+3}\\ x_{2i+1}&x_{2i+2}&x_{2i+4}\end{bmatrix}\right)
Proposition 2.13.

For n0n\geq 0, each graph GntG_{n}^{t}\in\mathcal{F} is chordal bipartite with vertex bipartition VrVcV_{r}\cup V_{c} of cardinalities

|Vr|=n2+2\displaystyle|V_{r}|=\left\lfloor\frac{n}{2}\right\rfloor+2
|Vc|=n2+2.\displaystyle|V_{c}|=\left\lceil\frac{n}{2}\right\rceil+2.
Proof.

We already know by Definition 2.5 that every graph GntG_{n}^{t} is bipartite for n0n\geq 0, with the bipartition above coming from the rows and columns of LntL_{n}^{t}. The cardinalities of the vertex sets follow from Definition 2.2. We prove the chordal bipartite property by induction on nn. It is clear for i=0i=0 and i=1i=1 that GifG_{i}^{f} is chordal bipartite for f𝔽2i+1f\in\mathbb{F}_{2}^{i+1}, since these graphs have fewer than six vertices. Now suppose GifG_{i}^{f} is chordal bipartite for ii with 1i<n0mod2(1mod2)1\leq i<n\equiv 0\mod 2(\equiv 1\mod 2), and consider GntG_{n}^{t} for t𝔽2n+1t\in\mathbb{F}_{2}^{n+1}. We know that the following array (or its transpose) is a subarray of LntL_{n}^{t} by Remark 2.12, and we include for reference the corresponding subgraph of GntG_{n}^{t} with vertices labeled by row and column.

[xj2nx2n+1xj2n+1x2n+2x2n+3x2n+4]\begin{bmatrix}x_{j_{2n}}&x_{2n+1}\\ x_{j_{2n+1}}&x_{2n+2}\\ x_{2n+3}&x_{2n+4}\end{bmatrix}
c1c1r1r1r2r2c2c2r3r3xj2nx_{j_{2n}}x2n+1x_{2n+1}xj2n+1x_{j_{2n+1}}x2n+2x_{2n+2}x2n+3x_{2n+3}x2n+4x_{2n+4}

We know the only difference between GntG_{n}^{t} and Gn1t^G_{n-1}^{\widehat{t}} is one vertex r3r_{3} corresponding to row n/2+2\left\lfloor{n/2}\right\rfloor+2 (column n/2+2\left\lceil{n/2}\right\rceil+2) and two edges x2n+3={r3,c1}x_{2n+3}=\{r_{3},c_{1}\} and x2n+4={r3,c2}x_{2n+4}=\{r_{3},c_{2}\}, where c1c_{1} corresponds to the column (row) containing x2n+3x_{2n+3} and c2c_{2} corresponds to column n/2+2\left\lceil{n/2}\right\rceil+2 (row n/2+2\left\lfloor{n/2}\right\rfloor+2). By Remark 2.2, degr3=2\deg r_{3}=2, since the only entries in row n/2+2\left\lfloor{n/2}\right\rfloor+2 (column n/2+2\left\lceil{n/2}\right\rceil+2) are x2n+3x_{2n+3} and x2n+4x_{2n+4}. Then any even cycle containing r3r_{3} must also contain x2n+3x_{2n+3} and x2n+4x_{2n+4}. Similarly, by the same remark, the only other edges with endpoint c2c_{2} are x2n+1x_{2n+1} and x2n+2x_{2n+2}, the entries added to make Ln1t^L_{n-1}^{\widehat{t}}, so we know that any even cycle containing x2n+4x_{2n+4} and x2n+3x_{2n+3} must contain either x2n+1x_{2n+1} or x2n+2x_{2n+2}. We see that any even cycle containing r3r_{3} and x2n+1x_{2n+1} is either a 4-cycle or has xj2nx_{j_{2n}} as a chord, and any even cycle containing r3r_{3} and x2n+2x_{2n+2} is either a 4-cycle or has xj2n+1x_{j_{2n+1}} as a chord. Thus every graph GntG_{n}^{t} is chordal bipartite for n0n\geq 0, with the bipartition above. ∎

Remark 2.14.

We note that the previous proposition ensures that our graphs are distinct from those studied in [8], [9], and [23], which are not chordal bipartite except for the first family in [8], in which every four-cycle shares exactly one edge with every other four-cycle (also distinct from our family except for the trivial case with only one four-cycle, corresponding to G0tG_{0}^{t}).

2.2. Toric Rings for \mathcal{F}

In this section, we develop the toric ring R(n,t)R(n,t) for each of the chordal bipartite graphs GntG_{n}^{t} in the family \mathcal{F}. We first show that the toric ideal IGntI_{G_{n}^{t}} of the graph GntG_{n}^{t} is the same as the ideal I(n,t)I(n,t) generated by the distinguished minors of LntL_{n}^{t}. We then demonstrate that for some nn and tt, these ideals do not arise from the join-meet ideals of lattices in a natural way, so that results in lattice theory do not apply to the general family \mathcal{F} in an obvious way.

We first define the toric ideal of a graph. For any graph GG with vertex set VV and edge set EE, there is a natural map π:k[E]k[V]\pi:k[E]\to k[V] taking an edge to the product of its endpoints. The polynomial subring in k[V]k[V] generated by the images of the edges under the map π\pi is denoted k[G]k[G] and is called the edge ring of GG. The kernel of π\pi is denoted IGI_{G} and is called the toric ideal of GG; the ring k[G]k[G] is isomorphic to k[E]/IGk[E]/I_{G}. In this work, we consider the toric ring k[E]/IGk[E]/I_{G} and the toric ideal IGI_{G} for our particular graphs. It is known in general that IGI_{G} is generated by binomial expressions coming from even closed walks in GG [27, Prop 3.1] and that the toric ideal of a chordal bipartite graph is generated by quadratic binomials coming from the 44-cycles of GG (see [20, Th 1.2]).

Let S(n)=k[x0,x2,x3,,x2n+4]S(n)=k[x_{0},x_{2},x_{3},\ldots,x_{2n+4}] be the polynomial ring in the edges of GntG_{n}^{t}. The edge ring for GntG_{n}^{t}\in\mathcal{F} is denoted by k[Gnt]k[G_{n}^{t}] and is isomorphic to the toric ring

R(n,t):=S(n)IGnt,R(n,t):=\frac{S(n)}{I_{G_{n}^{t}}},

where IGntI_{G_{n}^{t}} is the toric ideal of GntG_{n}^{t} [15, 5.3]. For the general construction of a toric ideal, we refer the reader to [24, Ch 4]. Our goal is to show that the toric ideal IGntI_{G_{n}^{t}} of GntG_{n}^{t} is equal to

I(n,t)=({distinguished minors of Lnt}).I(n,t)=(\{\text{distinguished minors of $L_{n}^{t}$}\}).
Proposition 2.15.

Let S(n)=k[x0,x2,x3,,x2n+4]S(n)=k[x_{0},x_{2},x_{3},\ldots,x_{2n+4}]. For GntG_{n}^{t}\in\mathcal{F}, we have

R(n,t)=S(n)I(n,t),R(n,t)=\frac{S(n)}{I(n,t)},

where

I(n,t)=({distinguished minors of Lnt}).I(n,t)=(\{\text{distinguished minors of $L_{n}^{t}$}\}).
Proof.

To prove this, we need only show that I(n,t)I(n,t) is the toric ideal IGntI_{G_{n}^{t}} of the graph GntG_{n}^{t}. By Definition 2.5, it is clear that the distinguished minors of LntL_{n}^{t} are in IGntI_{G_{n}^{t}}, corresponding to the 44-cycles of GntG_{n}^{t}. Since GG is chordal bipartite by Proposition 2.13, these are the only generators of IGntI_{G_{n}^{t}} [15, Cor 5.15]. ∎

Corollary 2.16.

The rings R(n,t)R(n,t) are normal Cohen-Macaulay rings.

Proof.

By Definition 2.5 and Proposition 2.15, the ring R(n,t)R(n,t) is the toric ring of a finite simple connected bipartite graph, and hence by Corollary 5.26 in [15], R(n,t)R(n,t) is Cohen-Macaulay for each nn and tt. The fact that each R(n,t)R(n,t) is normal follows from [25, Th 5.9, 7.1]. ∎

Because we know the distinguished minors of LntL_{n}^{t}, we are now able to characterize the generators for the toric ideal R(n,t)R(n,t) of GntG_{n}^{t}.

Remark 2.17.

By Proposition 2.9, the generators s1,,s2n+1s_{1},\ldots,s_{2n+1} for IGntI_{G_{n}^{t}} may be summarized as follows. For integers ii such that 1in1\leq i\leq n, set

s1\displaystyle s_{1} =\displaystyle= x2x3xj1x4\displaystyle x_{2}x_{3}-x_{j_{1}}x_{4}
s2i\displaystyle s_{2i} =\displaystyle= x2i+1x2i+3xj2ix2i+4\displaystyle x_{2i+1}x_{2i+3}-x_{j_{2i}}x_{2i+4}
s2i+1\displaystyle s_{2i+1} =\displaystyle= x2i+2x2i+3xj2i+1x2i+4,\displaystyle x_{2i+2}x_{2i+3}-x_{j_{2i+1}}x_{2i+4},

where the nonnegative integers jkj_{k} are as in Remark 2.11, that is, j1=j2=0j_{1}=j_{2}=0, j3=2j_{3}=2, and for i2i\geq 2, we have

ti+1\displaystyle t_{i+1} =\displaystyle= 0j2i=j2i2j2i+1=2i1\displaystyle 0\iff j_{2i}=j_{2i-2}\iff j_{2i+1}=2i-1
ti+1\displaystyle t_{i+1} =\displaystyle= 1j2i=j2i1j2i+1=2i.\displaystyle 1\iff j_{2i}=j_{2i-1}\iff j_{2i+1}=2i.

We note that the number of generators depends on nn and that the jkj_{k} depend on tt, but we may ignore dependence on tt when working with general jkj_{k}. We sometimes call s1,,s2n+1s_{1},\ldots,s_{2n+1} the standard generators of IGntI_{G_{n}^{t}}, and show in Section 2.3 that for certain nn and tt, they are not equal to the usual generators for the join-meet ideal of any lattice DD.

Example 2.18.

We consider the toric ideal of a graph in 1\mathcal{F}_{1}. For n=5n=5 and t=(1,1,,1)t=(1,1,\ldots,1), by Remark 2.11 we have j1=j2=0j_{1}=j_{2}=0, j3=2j_{3}=2, j2i=j2i1j_{2i}=j_{2i-1} and j2i+1=2ij_{2i+1}=2i for i2i\geq 2, so that

R(5,(1,1,,1))=k[x0,x2,,x14]IG5(1,1,,1),R(5,(1,1,\ldots,1))=\frac{k[x_{0},x_{2},\ldots,x_{14}]}{I_{G_{5}^{(1,1,\ldots,1)}}},

where IG5(1,1,,1)I_{G_{5}^{(1,1,\ldots,1)}} is generated by the distinguished minors of L5(1,1,,1)L_{5}^{(1,1,\ldots,1)}:

s1\displaystyle s_{1} =x2x3x0x4\displaystyle=x_{2}x_{3}-x_{0}x_{4} s2\displaystyle s_{2} =x3x5x0x6\displaystyle=x_{3}x_{5}-x_{0}x_{6}
s3\displaystyle s_{3} =x4x5x2x6\displaystyle=x_{4}x_{5}-x_{2}x_{6} s4\displaystyle s_{4} =x5x7x2x8\displaystyle=x_{5}x_{7}-x_{2}x_{8}
s5\displaystyle s_{5} =x6x7x4x8\displaystyle=x_{6}x_{7}-x_{4}x_{8} s6\displaystyle s_{6} =x7x9x4x10\displaystyle=x_{7}x_{9}-x_{4}x_{10}
s7\displaystyle s_{7} =x8x9x6x10\displaystyle=x_{8}x_{9}-x_{6}x_{10} s8\displaystyle s_{8} =x9x11x6x12\displaystyle=x_{9}x_{11}-x_{6}x_{12}
s9\displaystyle s_{9} =x10x11x8x12\displaystyle=x_{10}x_{11}-x_{8}x_{12} s10\displaystyle s_{10} =x11x13x8x14\displaystyle=x_{11}x_{13}-x_{8}x_{14}
s11\displaystyle s_{11} =x12x13x10x14.\displaystyle=x_{12}x_{13}-x_{10}x_{14}.

2.3. Distinction From Join-Meet Ideals of Lattices

We saw in Example 2.4 and Proposition 2.15 that if Gnt1G_{n}^{t}\in\mathcal{F}_{1}\subset\mathcal{F}, then IGntI_{G_{n}^{t}} is a ladder determinantal ideal for n2n\geq 2. It is known that a ladder determinantal ideal is equal to the join-meet ideal of a (distributive) lattice (indeed, with a natural partial ordering which decreases along rows and columns of LntL_{n}^{t} we obtain such a lattice). Some algebraic information such as regularity and projective dimension may be easily derived for some join-meet ideals of distributive lattices (see, for example, Chapter 6 of [15]). We spend some time in this section establishing that not all rings R(n,t)R(n,t)\in\mathcal{F} arise from a lattice in a natural way (see Remark 2.20), and so there does not seem to be any obvious way to obtain our results in Section 3 from the literature on join-meet ideals of distributive lattices or on ladder determinantal ideals. The results in Section 3 may be viewed as an extension of what may already be derived for the family 1\mathcal{F}_{1} from the existing literature.

The following five lemmas serve to provide machinery to show that there is at least one ring in the family \mathcal{F}, namely R(5,(1,1,1,1,1,0))R(5,(1,1,1,1,1,0)), whose toric ideal does not come from a lattice on the set {x0,x2,,x14}\{x_{0},x_{2},\ldots,x_{14}\} in any obvious way. That is, we show that the standard generators of IG5tI_{G_{5}^{t}}, the sks_{k} from Remark 2.17, are not equal to the standard generators (see Definition 2.19) for any lattice DD on {x0,x2,,x14}\{x_{0},x_{2},\ldots,x_{14}\}.

Before we begin, we introduce some definitions and notation that we use extensively throughout.

Definition 2.19.

The join-meet ideal of a lattice is defined from the join (least upper bound) xyx\vee y and meet (greatest lower bound) xyx\wedge y of each pair of incomparable elements x,yLx,y\in L. In this work, a standard generator of the join-meet ideal of a lattice DD is a nonzero element of one of the following four forms:

xaxb(xaxb)(xaxb)\displaystyle x_{a}x_{b}-(x_{a}\vee x_{b})(x_{a}\wedge x_{b}) =\displaystyle= xaxb(xaxb)(xaxb)\displaystyle x_{a}x_{b}-(x_{a}\wedge x_{b})(x_{a}\vee x_{b})
(xaxb)(xaxb)xaxb\displaystyle(x_{a}\vee x_{b})(x_{a}\wedge x_{b})-x_{a}x_{b} =\displaystyle= (xaxb)(xaxb)xaxb\displaystyle(x_{a}\wedge x_{b})(x_{a}\vee x_{b})-x_{a}x_{b}

for xa,xbLx_{a},x_{b}\in L. We sometimes refer to such an element as a standard generator of DD (the join-meet ideal is defined by analogous generators in the literature, though sometimes a,bLa,b\in L instead of xax_{a} and xbx_{b}). We note that for a standard generator, the pair {xa,xb}\{x_{a},x_{b}\} is an incomparable pair, and the pair {(xaxb),(xaxb)}\{(x_{a}\vee x_{b}),(x_{a}\wedge x_{b})\} is a comparable pair.

Though we are in a commutative ring, we provide all possible orderings for factors within the terms of a standard generator to emphasize that either factor of the monomial

(xaxb)(xaxb)=(xaxb)(xaxb)(x_{a}\vee x_{b})(x_{a}\wedge x_{b})=(x_{a}\wedge x_{b})(x_{a}\vee x_{b})

may be the join or the meet of xax_{a} and xbx_{b}.

Remark 2.20.

We give an explanation for why it makes sense to focus only on the standard generators of a join-meet ideal. We recall that the standard generators sks_{k} for IGntI_{G_{n}^{t}} from Remark 2.17 come from distinct 2×22\times 2 arrays within the ladder-like structure LntL_{n}^{t} and recognize that either monomial of sks_{k} determines its 2×22\times 2 array. Then an element of the form abcdab-cd in IG5tI_{G_{5}^{t}} with a,b,c,d{x0,x2,x3,,x14}a,b,c,d\in\{x_{0},x_{2},x_{3},\ldots,x_{14}\} must be equal to ±si\pm s_{i} for some ii, since a nontrivial sum of sks_{k} with coefficients in {1,1}\{-1,1\} either has more than two terms or is equal to sis_{i} for some ii, and other coefficients would be extraneous. Then any generating set for IG5tI_{G_{5}^{t}} where each element has the form abcdab-cd in IG5tI_{G_{5}^{t}} with a,b,c,d{x0,x2,x3,,x14}a,b,c,d\in\{x_{0},x_{2},x_{3},\ldots,x_{14}\} must consist of all the sks_{k} (up to sign). We conclude that it is natural to check whether the sks_{k} are standard generators of a lattice DD, instead of non-standard generators.

Definition 2.21.

Given a standard generator s=uzwvs=uz-wv of a lattice DD, where u,v,w,zDu,v,w,z\in D, let Fs𝔽2F_{s}\in\mathbb{F}_{2} be defined as follows:

  • If Fs=0F_{s}=0, the elements in the first (positive) monomial of ss are not comparable in DD (so the elements in the second (negative) monomial of ss are comparable in DD).

  • If Fs=1F_{s}=1, the elements in the negative monomial of ss are not comparable in DD (so the elements in the positive monomial of ss are comparable in DD).

For a given list s1,s2,,sms_{1},s_{2},\ldots,s_{m} of standard generators of a lattice DD, we use

F=(F1,,Fm)𝔽2m,F=(F_{1},\ldots,F_{m})\in\mathbb{F}_{2}^{m},

where Fj=FsjF_{j}=F_{s_{j}}, to encode the comparability of the variables in these generators.

We note that exactly one of Fj=0F_{j}=0 or Fj=1F_{j}=1 happens for each jj; we are merely encoding which monomial in sjs_{j} corresponds to xaxbx_{a}x_{b}, and which to (xaxb)(xaxb)=(xaxb)(xaxb)(x_{a}\vee x_{b})(x_{a}\wedge x_{b})=(x_{a}\wedge x_{b})(x_{a}\vee x_{b}).

Notation 2.22.

We use the notation u>{w,v}u>\{w,v\} if u>wu>w and u>vu>v in a lattice DD, and {w,v}>z\{w,v\}>z if w>zw>z and v>zv>z in DD.

In the first lemma, we begin by showing what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the following array as standard generators:

abecdf\begin{matrix}a&b&e\\ c&d&f\end{matrix}

Lemma 2.23.

Suppose

s1\displaystyle s_{1} =\displaystyle= bcad\displaystyle bc-ad
s2\displaystyle s_{2} =\displaystyle= ceaf\displaystyle ce-af
s3\displaystyle s_{3} =\displaystyle= debf\displaystyle de-bf

are standard generators of a lattice DD. Let F𝔽23F\in\mathbb{F}_{2}^{3} be defined for these three elements as in Definition 2.21. Then up to relabeling of variables,

F{{0,0,0},{0,0,1},{0,1,1}}.F\in\{\{0,0,0\},\{0,0,1\},\{0,1,1\}\}.
Proof.

We first note that some of the cases we consider are equivalent. If we relabel variables according to the permutation (ac)(bd)(ef)(ac)(bd)(ef), we see that

F={i,j,k}{1i,1j,1k}.F=\{i,j,k\}\equiv\{1-i,1-j,1-k\}.

This limits the cases we need to consider to

F{{0,0,0},{0,0,1},{0,1,0},{0,1,1}}.F\in\{\{0,0,0\},\{0,0,1\},\{0,1,0\},\{0,1,1\}\}.

That is, we only need to show that the case F={0,1,0}F=\{0,1,0\} is impossible.

Let F1=0F_{1}=0. Then without loss of generality, up to reversing the order in the lattice (which does not affect the join-meet ideal), we have a>{b,c}>da>\{b,c\}>d. If F2=1F_{2}=1, we have e>{a,f}>ce>\{a,f\}>c, so e>{b,f}>de>\{b,f\}>d and hence F3=1F_{3}=1. We conclude that the case F={0,1,0}F=\{0,1,0\} is impossible. ∎

In the second lemma, we show what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the ladder

abecdfgh\begin{matrix}a&b&e\\ c&d&f\\ &g&h\\ \end{matrix}

as standard generators, and which meets certain comparability conditions.

Lemma 2.24.

Suppose

s1\displaystyle s_{1} =\displaystyle= bcad\displaystyle bc-ad
s2\displaystyle s_{2} =\displaystyle= ceaf\displaystyle ce-af
s3\displaystyle s_{3} =\displaystyle= debf\displaystyle de-bf
s4\displaystyle s_{4} =\displaystyle= egbh\displaystyle eg-bh
s5\displaystyle s_{5} =\displaystyle= fgdh\displaystyle fg-dh

are standard generators of a lattice DD, and that {a,g}\{a,g\},{a,h}\{a,h\},{c,g}\{c,g\}, and {c,h}\{c,h\} are comparable pairs in DD. Let F𝔽25F\in\mathbb{F}_{2}^{5} be defined for these five elements as in 2.21. Then up to relabeling of variables, F={0,0,0,0,0}F=\{0,0,0,0,0\}.

Proof.

We first note that with natural relabeling, both {s1,s2,s3}\{s_{1},s_{2},s_{3}\} and {s3,s4,s5}\{s_{3},s_{4},s_{5}\} satisfy the hypotheses of Lemma 2.23, so if we let FF be defined as in Definition 2.21, this limits the cases we need to consider to 5-tuples whose first three elements and whose last three elements satisfy the conclusion of Lemma 2.23. We note that some of the cases we consider are equivalent. If we relabel variables according to the permutation (ac)(bd)(ef)(ac)(bd)(ef), we see that F={i,j,k,l,m}{1i,1j,1k,m,l}F=\{i,j,k,l,m\}\equiv\{1-i,1-j,1-k,m,l\}, and if we relabel the variables according to the permutation (be)(df)(gh)(be)(df)(gh), we have F={i,j,k,l,m}{j,i,1k,1l,1m}F=\{i,j,k,l,m\}\equiv\{j,i,1-k,1-l,1-m\}. The permutation (ah)(cg)(bf)(ah)(cg)(bf) yields F={i,j,k,l,m}{m,l,k,j,i}F=\{i,j,k,l,m\}\equiv\{m,l,k,j,i\}. Then by Lemma 2.23 we have the eighteen cases

{0,0,0,0,0}{1,1,1,0,0}{1,1,0,1,1}{0,0,1,1,1}\{0,0,0,0,0\}\equiv\{1,1,1,0,0\}\equiv\{1,1,0,1,1\}\equiv\{0,0,1,1,1\}
{0,0,0,0,1}{1,1,1,1,0}{1,1,0,0,1}{0,0,1,1,0}\displaystyle\{0,0,0,0,1\}\equiv\{1,1,1,1,0\}\equiv\{1,1,0,0,1\}\equiv\{0,0,1,1,0\} \displaystyle\equiv {0,1,1,0,0}\displaystyle\{0,1,1,0,0\}
\displaystyle\equiv {1,0,0,0,0}\displaystyle\{1,0,0,0,0\}
\displaystyle\equiv {0,1,1,1,1}\displaystyle\{0,1,1,1,1\}
\displaystyle\equiv {1,0,0,1,1}\displaystyle\{1,0,0,1,1\}
{0,0,0,1,1}{1,1,1,1,1}{1,1,0,0,0}{0,0,1,0,0}\{0,0,0,1,1\}\equiv\{1,1,1,1,1\}\equiv\{1,1,0,0,0\}\equiv\{0,0,1,0,0\}
{0,1,1,1,0}{1,0,0,0,1}\{0,1,1,1,0\}\equiv\{1,0,0,0,1\}

Case 1: F={0,0,0,0,1}F=\{0,0,0,0,1\}. Since F1=0F_{1}=0, without loss of generality (reversing the order on the entire lattice if needed) we have a>{b,c}>da>\{b,c\}>d. Then F2=F3=F4=0F_{2}=F_{3}=F_{4}=0 and F5=1F_{5}=1, with the ordering chosen, yield

a>{c,e}>f\displaystyle a>\{c,e\}>f
b>{d,e}>f\displaystyle b>\{d,e\}>f
b>{e,g}>h\displaystyle b>\{e,g\}>h
g>{d,h}>f.\displaystyle g>\{d,h\}>f.

If c>gc>g, then a>{b,c}>g>da>\{b,c\}>g>d, but then bcadbc-ad is not a standard generator of DD, and this is a contradiction. If c<gc<g, then c<g<bc<g<b so that both {b,c}\{b,c\} and {a,d}\{a,d\} from s1s_{1} are comparable pairs, but this is a contradiction. We conclude that the case {0,0,0,0,1}\{0,0,0,0,1\} is impossible.

Because of the comparability of the pair {c,g}\{c,g\}, the case F={0,1,1,1,0}F=\{0,1,1,1,0\} forces comparability of {f,g}\{f,g\} or {b,c}\{b,c\} and hence yields a contradiction, as the reader may verify. In the case F={0,0,0,1,1}F=\{0,0,0,1,1\}, the comparability of {c,g}\{c,g\} forces either the comparability of {b,c}\{b,c\} (a contradiction), or d<{b,c}<a<gd<\{b,c\}<a<g, up to reversing the order in the lattice, since s1s_{1} is a standard generator. Since {a,h}\{a,h\} is a comparable pair, it immediately follows that either {b,h}\{b,h\} is comparable (a contradiction) or e<{b,h}<a<ge<\{b,h\}<a<g, which is a contradiction since s4s_{4} is a standard generator, as the reader may verify. These cases are compatible with the given relabelings and thus conclude our proof. ∎

In the third lemma, we show what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the ladder

abecdfighj\begin{matrix}a&b&e&\\ c&d&f&i\\ &g&h&j\end{matrix}

as standard generators and which meets certain comparability conditions.

Lemma 2.25.

Suppose

s1\displaystyle s_{1} =\displaystyle= bcad\displaystyle bc-ad
s2\displaystyle s_{2} =\displaystyle= ceaf\displaystyle ce-af
s3\displaystyle s_{3} =\displaystyle= debf\displaystyle de-bf
s4\displaystyle s_{4} =\displaystyle= egbh\displaystyle eg-bh
s5\displaystyle s_{5} =\displaystyle= fgdh\displaystyle fg-dh
s6\displaystyle s_{6} =\displaystyle= gidj\displaystyle gi-dj
s7\displaystyle s_{7} =\displaystyle= hifj\displaystyle hi-fj

are standard generators of a lattice DD, and that {a,g}\{a,g\}, {a,h}\{a,h\}, {c,g}\{c,g\}, {c,h}\{c,h\}, {b,i}\{b,i\}, {b,j}\{b,j\}, {e,i}\{e,i\}, and {e,j}\{e,j\} are comparable pairs in DD. Let F𝔽27F\in\mathbb{F}_{2}^{7} be defined for these seven elements as in Definition 2.21. Then up to relabeling of variables, F={0,0,0,0,0,0,0}F=\{0,0,0,0,0,0,0\}.

Proof.

We first note that with natural relabeling, both {s1,s2,s3,s4,s5}\{s_{1},s_{2},s_{3},s_{4},s_{5}\} and {s3,s4,s5,s6,s7}\{s_{3},s_{4},s_{5},s_{6},s_{7}\} satisfy the hypotheses of Lemma 2.24, so if we let FF be defined as in Definition 2.21, this limits the cases we need to consider to 7-tuples whose first five elements and whose last five elements satisfy the conclusion of Lemma 2.24. The only possible cases are {0,0,0,0,0,0,0}\{0,0,0,0,0,0,0\} and {0,0,1,1,1,0,0}\{0,0,1,1,1,0,0\}. If we relabel variables according to the permutation (be)(df)(gh)(be)(df)(gh), we see that F={i,j,k,l,m,n,o}{j,i,1k,1l,1m,o,n}F=\{i,j,k,l,m,n,o\}\equiv\{j,i,1-k,1-l,1-m,o,n\}, so that these two cases are equivalent. Then up to relabeling of variables, F={0,0,0,0,0,0,0}F=\{0,0,0,0,0,0,0\}. ∎

In the fourth lemma, we show what restrictions we must have on a lattice whose join-meet ideal contains the 2-minors of the ladder

abecdfighjkl\begin{matrix}a&b&e&\\ c&d&f&i\\ &g&h&j\\ &&k&l\end{matrix}

as standard generators, and which meets certain comparability conditions.

Lemma 2.26.

Suppose

s1\displaystyle s_{1} =\displaystyle= bcad\displaystyle bc-ad
s2\displaystyle s_{2} =\displaystyle= ceaf\displaystyle ce-af
s3\displaystyle s_{3} =\displaystyle= debf\displaystyle de-bf
s4\displaystyle s_{4} =\displaystyle= egbh\displaystyle eg-bh
s5\displaystyle s_{5} =\displaystyle= fgdh\displaystyle fg-dh
s6\displaystyle s_{6} =\displaystyle= gidj\displaystyle gi-dj
s7\displaystyle s_{7} =\displaystyle= hifj\displaystyle hi-fj
s8\displaystyle s_{8} =\displaystyle= ikfl\displaystyle ik-fl
s9\displaystyle s_{9} =\displaystyle= jkhl\displaystyle jk-hl

are standard generators of a lattice DD, and that {a,g}\{a,g\}, {a,h}\{a,h\}, {c,g}\{c,g\}, {c,h}\{c,h\}, {b,i}\{b,i\}, {b,j}\{b,j\}, {e,i}\{e,i\}, {e,j}\{e,j\}, {d,k}\{d,k\}, {d,l}\{d,l\}, {g,k}\{g,k\}, and {g,l}\{g,l\} are comparable pairs in DD. Let F𝔽29F\in\mathbb{F}_{2}^{9} be defined for these nine elements as in Definition 2.21. Then F={0,0,0,0,0,0,0,0,0}F=\{0,0,0,0,0,0,0,0,0\}.

Proof.

We first note that with natural relabeling, both {s1,s2,s3,s4,s5,s6,s7}\{s_{1},s_{2},s_{3},s_{4},s_{5},s_{6},s_{7}\} and
{s3,s4,s5,s6,s7,s8,s9}\{s_{3},s_{4},s_{5},s_{6},s_{7},s_{8},s_{9}\} satisfy the hypotheses of Lemma 2.25, so if we let FF be defined as in Definition 2.21, this limits the cases we need to consider to 9-tuples whose first seven entries and whose last seven entries satisfy the conclusion of Lemma 2.25. We see by Lemma 2.25 that F={0,0,0,0,0,0,0,0,0}F=\{0,0,0,0,0,0,0,0,0\}. ∎

We now have the machinery necessary to show that for t=(1,1,1,1,1,0)t=(1,1,1,1,1,0), IG5tI_{G_{5}^{t}} does not come from a lattice. In our proof, we use the previous four lemmas and the fact that IG5tI_{G_{5}^{t}} is generated by the distinguished minors of L5(1,1,1,1,1,0)L_{5}^{(1,1,1,1,1,0)}:

x0x2x5x3x4x6x9x13x7x8x10x11x12x14\begin{matrix}x_{0}&x_{2}&x_{5}&&\\ x_{3}&x_{4}&x_{6}&x_{9}&x_{13}\\ &x_{7}&x_{8}&x_{10}&\\ &&x_{11}&x_{12}&x_{14}\end{matrix}

Proposition 2.27.

Let n=5n=5 and t=(1,1,1,1,1,0)t=(1,1,1,1,1,0). Then the set of standard generators for IG5tI_{G_{5}^{t}} is not equal to the complete set of standard generators (up to sign) of any (classical) lattice.

Proof.

By Remark 2.17 and choice of n=5n=5 and t=(1,1,1,1,1,0)t=(1,1,1,1,1,0), the generators of IG5tI_{G_{5}^{t}} are

s1\displaystyle s_{1} =x2x3x0x4\displaystyle=x_{2}x_{3}-x_{0}x_{4} s2\displaystyle s_{2} =x3x5x0x6\displaystyle=x_{3}x_{5}-x_{0}x_{6}
s3\displaystyle s_{3} =x4x5x2x6\displaystyle=x_{4}x_{5}-x_{2}x_{6} s4\displaystyle s_{4} =x5x7x2x8\displaystyle=x_{5}x_{7}-x_{2}x_{8}
s5\displaystyle s_{5} =x6x7x4x8\displaystyle=x_{6}x_{7}-x_{4}x_{8} s6\displaystyle s_{6} =x7x9x4x10\displaystyle=x_{7}x_{9}-x_{4}x_{10}
s7\displaystyle s_{7} =x8x9x6x10\displaystyle=x_{8}x_{9}-x_{6}x_{10} s8\displaystyle s_{8} =x9x11x6x12\displaystyle=x_{9}x_{11}-x_{6}x_{12}
s9\displaystyle s_{9} =x10x11x8x12\displaystyle=x_{10}x_{11}-x_{8}x_{12} s10\displaystyle s_{10} =x11x13x6x14\displaystyle=x_{11}x_{13}-x_{6}x_{14}
s11\displaystyle s_{11} =x12x13x9x14\displaystyle=x_{12}x_{13}-x_{9}x_{14}

Suppose a lattice DD exists whose complete set of standard generators (up to sign) equals {s1,,s11s_{1},\ldots,s_{11}}. We note that if the monomial xixjx_{i}x_{j} does not appear in any of the sks_{k}, then {xi,xj}\{x_{i},x_{j}\} is a comparable pair, since otherwise ±(xixj(xixj)(xixj))\pm(x_{i}x_{j}-(x_{i}\vee x_{j})(x_{i}\wedge x_{j})) would be in the set of standard generators of DD. Thus the pairs {x0,x7}\{x_{0},x_{7}\}, {x0,x8}\{x_{0},x_{8}\}, {x3,x7}\{x_{3},x_{7}\}, {x3,x8}\{x_{3},x_{8}\}, {x2,x9}\{x_{2},x_{9}\}, {x2,x10}\{x_{2},x_{10}\}, {x5,x9}\{x_{5},x_{9}\}, {x5,x10}\{x_{5},x_{10}\}, {x4,x11}\{x_{4},x_{11}\}, {x4,x12}\{x_{4},x_{12}\}, {x7,x11}\{x_{7},x_{11}\}, {x7,x12}\{x_{7},x_{12}\}, {x10,x13}\{x_{10},x_{13}\}, and {x10,x14}\{x_{10},x_{14}\} are comparable pairs in DD. Let F𝔽211F\in\mathbb{F}_{2}^{11} be defined as in Definition 2.21. Then with natural relabeling of the first nine relations, this lattice satisfies the hypotheses of Lemma 2.26, so the only cases we need to consider are F={0,0,0,0,0,0,0,0,0,a,b}F=\{0,0,0,0,0,0,0,0,0,a,b\}.

Since F1=0F_{1}=0, without loss of generality, we have x0>{x2,x3}>x4x_{0}>\{x_{2},x_{3}\}>x_{4}. Then with the ordering chosen, the fact that F3=F5=F7=F9=0F_{3}=F_{5}=F_{7}=F_{9}=0 yields

x2>{x4,x5}>x6\displaystyle x_{2}>\{x_{4},x_{5}\}>x_{6}
x4>{x6,x7}>x8\displaystyle x_{4}>\{x_{6},x_{7}\}>x_{8}
x6>{x8,x9}>x10\displaystyle x_{6}>\{x_{8},x_{9}\}>x_{10}
x8>{x10,x11}>x12.\displaystyle x_{8}>\{x_{10},x_{11}\}>x_{12}.

The reader may verify that b=0b=0 and b=1b=1 both yield contradictions based on inspecting the standard generator s11s_{11} in light of the comparability of the pairs {x10,x13}\{x_{10},x_{13}\} and {x10,x14}\{x_{10},x_{14}\}, respectively, using the same technique employed in Case 1 of the proof of Lemma 2.24.

We conclude that there is no lattice whose complete set of standard generators (up to sign) equals the set of standard generators of IG5(1,1,1,1,1,0)I_{G_{5}^{(1,1,1,1,1,0)}}. ∎

3. Properties of the Family of Toric Rings

In Section 2, we defined a family of toric rings, the rings R(n,t)R(n,t) coming from the family \mathcal{F}, and we demonstrated some context for these rings in the area of graph theory. Now we investigate some of the algebraic properties of each R(n,t)R(n,t). We develop proofs to establish dimension, regularity, and multiplicity.

3.1. Dimension and System of Parameters

We use the the degree reverse lexicographic monomial order with x0>x2>x3>x_{0}>x_{2}>x_{3}>\cdots throughout this section, and denote it by >>. We show that the standard generators sks_{k} given in Remark 2.17 are a Gröbner basis for IGntI_{G_{n}^{t}} with respect to >>.

Lemma 3.1.

If s1,,s2n+1s_{1},\ldots,s_{2n+1} are as in Remark 2.17, then B={s1,,s2n+1}B=\{s_{1},\ldots,s_{2n+1}\} is a Gröbner basis for IGntI_{G_{n}^{t}} with respect to >>.

Proof.

This is a straightforward computation using Buchberger’s Criterion and properties of the jkj_{k} from Remark 2.11. By Remark 2.17, for 1in1\leq i\leq n the ideal IGntI_{G_{n}^{t}} is generated by

s1\displaystyle s_{1} =\displaystyle= x2x3xj1x4\displaystyle x_{2}x_{3}-x_{j_{1}}x_{4}
s2i\displaystyle s_{2i} =\displaystyle= x2i+1x2i+3xj2ix2i+4\displaystyle x_{2i+1}x_{2i+3}-x_{j_{2i}}x_{2i+4}
s2i+1\displaystyle s_{2i+1} =\displaystyle= x2i+2x2i+3xj2i+1x2i+4,\displaystyle x_{2i+2}x_{2i+3}-x_{j_{2i+1}}x_{2i+4},

where j1=j2=0j_{1}=j_{2}=0, j3=2j_{3}=2, and for i2i\geq 2, we have

ti+1\displaystyle t_{i+1} =\displaystyle= 0j2i=j2i2j2i+1=2i1\displaystyle 0\iff j_{2i}=j_{2i-2}\iff j_{2i+1}=2i-1
ti+1\displaystyle t_{i+1} =\displaystyle= 1j2i=j2i1j2i+1=2i.\displaystyle 1\iff j_{2i}=j_{2i-1}\iff j_{2i+1}=2i.

If we adopt SS-polynomial notation Si,jS_{i,j} for the SS-polynomial of sis_{i} and sjs_{j}, then the cases to consider are

S2i1,2i and S2i,2i+1 for 1inS_{2i-1,2i}\text{ and }S_{2i,2i+1}\text{ for }1\leq i\leq n
S2i,2i+2 for 1in1.S_{2i,2i+2}\text{ for }1\leq i\leq n-1.

To give a flavor of the computation involved, we show the case S2i1,2iS_{2i-1,2i} for 1in1\leq i\leq n, and leave the remaining cases to the reader. We show in each subcase that S2i1,2iS_{2i-1,2i} is equal to a sum of basis elements with coefficients in S(n)S(n), so that the reduced form of S2i1,2iS_{2i-1,2i} is zero in each subcase. We have

S2i1,2i\displaystyle S_{2i-1,2i} =\displaystyle= x2i+3(x2ix2i+1xj2i1x2i+2)x2i(x2i+1x2i+3xj2ix2i+4)\displaystyle x_{2i+3}(x_{2i}x_{2i+1}-x_{j_{2i-1}}x_{2i+2})-x_{2i}(x_{2i+1}x_{2i+3}-x_{j_{2i}}x_{2i+4})
=\displaystyle= xj2i1x2i+2x2i+3+xj2ix2ix2i+4\displaystyle-x_{j_{2i-1}}x_{2i+2}x_{2i+3}+x_{j_{2i}}x_{2i}x_{2i+4}

Case 1: If i2i\geq 2 and ti+1=0t_{i+1}=0, then j2i=j2i2j_{2i}=j_{2i-2} and j2i+1=2i1j_{2i+1}=2i-1, so we have

S2i1,2i\displaystyle S_{2i-1,2i} =\displaystyle= xj2i1x2i+2x2i+3+xj2i2x2ix2i+4\displaystyle-x_{j_{2i-1}}x_{2i+2}x_{2i+3}+x_{j_{2i-2}}x_{2i}x_{2i+4}

Case 1.1: If in addition i3i\geq 3 and ti=0t_{i}=0, then j2i2=j2i4j_{2i-2}=j_{2i-4} and j2i1=2i3j_{2i-1}=2i-3, so we have

S2i1,2i\displaystyle S_{2i-1,2i} =\displaystyle= x2i3x2i+2x2i+3+xj2i4x2ix2i+4\displaystyle-x_{2i-3}x_{2i+2}x_{2i+3}+x_{j_{2i-4}}x_{2i}x_{2i+4}
=\displaystyle= x2i3(x2i+2x2i+3xj2i+1x2i+4)+x2i+4(x2i3x2i1xj2i4x2i)\displaystyle-x_{2i-3}(x_{2i+2}x_{2i+3}-x_{j_{2i+1}}x_{2i+4})+x_{2i+4}(x_{2i-3}x_{2i-1}-x_{j_{2i-4}}x_{2i})
=\displaystyle= x2i3s2i+1+x2i+4s2i4.\displaystyle-x_{2i-3}s_{2i+1}+x_{2i+4}s_{2i-4}.

Case 1.2: If in addition i=2i=2 or i3i\geq 3 and ti=1t_{i}=1, then j2i2=j2i3j_{2i-2}=j_{2i-3} and j2i1=2i2j_{2i-1}=2i-2, so we have

S2i1,2i\displaystyle S_{2i-1,2i} =\displaystyle= x2i2x2i+2x2i+3+xj2i3x2ix2i+4\displaystyle-x_{2i-2}x_{2i+2}x_{2i+3}+x_{j_{2i-3}}x_{2i}x_{2i+4}
=\displaystyle= x2i2(x2i+2x2i+3xj2i+1x2i+4)+x2i+4(x2i2x2i1xj2i3x2i)\displaystyle-x_{2i-2}(x_{2i+2}x_{2i+3}-x_{j_{2i+1}}x_{2i+4})+x_{2i+4}(x_{2i-2}x_{2i-1}-x_{j_{2i-3}}x_{2i})
=\displaystyle= x2i2s2i+1+x2i+4s2i3.\displaystyle-x_{2i-2}s_{2i+1}+x_{2i+4}s_{2i-3}.

Case 2: If i=1i=1 or if i2i\geq 2 and ti+1=1t_{i+1}=1, then j2i=j2i1j_{2i}=j_{2i-1} and j2i+1=2ij_{2i+1}=2i, so we have

S2i1,2i\displaystyle S_{2i-1,2i} =\displaystyle= xj2ix2i+2x2i+3+xj2ix2ix2i+4\displaystyle-x_{j_{2i}}x_{2i+2}x_{2i+3}+x_{j_{2i}}x_{2i}x_{2i+4}
=\displaystyle= xj2is2i+1.\displaystyle-x_{j_{2i}}s_{2i+1}.

This concludes the case S2i1,2iS_{2i-1,2i} for 1in11\leq i\leq n-1. The remaining cases are similar.

Corollary 3.2.

The ring R(n,t)R(n,t) is Koszul for all nn and all tt.

Proof.

Since IGntI_{G_{n}^{t}} has a quadratic Gröbner basis, the ring R(n,t)R(n,t) is Koszul for all nn and all tt due to [15, Th 2.28]. ∎

Corollary 3.3.

The initial ideal for IGntI_{G_{n}^{t}} with respect to the degree reverse lexicographic monomial order >> is

in>IGnt=(x2x3,{x2i+1x2i+3,x2i+2x2i+31in}).\text{in}_{>}I_{G_{n}^{t}}=(x_{2}x_{3},\{x_{2i+1}x_{2i+3},x_{2i+2}x_{2i+3}\mid 1\leq i\leq n\}).

We note that in>IGnt\text{in}_{>}I_{G_{n}^{t}} does not depend on tt, which will be useful for the following sections.

We use the initial ideal in>IGnt\text{in}_{>}I_{G_{n}^{t}} from Corollary 3.3 and direct computation to show the Krull dimension of R(n,t)R(n,t). As a corollary, we obtain the projective dimension of R(n,t)R(n,t). We note that the Krull dimension, like the initial ideal, does not depend on tt. We refer the reader to Remark 2.17 for a reminder of how to think of the toric ring

R(n,t)=S(n)IGnt=k[x0,x2,x3,,x2n+4]IGntR(n,t)=\frac{S(n)}{I_{G_{n}^{t}}}=\frac{k[x_{0},x_{2},x_{3},\ldots,x_{2n+4}]}{I_{G_{n}^{t}}}

in the context of this work.

Theorem 3.4.

The Krull dimension of R(n,t)R(n,t) is

dimR(n,t)=n+3.\dim R(n,t)=n+3.
Proof.

Let >> be the degree reverse lexicographic monomial order with

x0>x2>x3>>x2n+4.x_{0}>x_{2}>x_{3}>\cdots>x_{2n+4}.

By Corollary 3.3, the initial ideal of IGntI_{G_{n}^{t}} with respect to >> is

in>IGnt=(x2x3,{x2i+1x2i+3,x2i+2x2i+31in}).\text{in}_{>}I_{G_{n}^{t}}=(x_{2}x_{3},\{x_{2i+1}x_{2i+3},x_{2i+2}x_{2i+3}\mid 1\leq i\leq n\}).

Since S(n)/(in>IGnt)S(n)/(\text{in}_{>}I_{G_{n}^{t}}) and R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}} are known to have the same Krull dimension (see for example [6, Props 9.3.4 and 9.3.12]), it suffices to prove that

dimS(n)/(in>IGnt)=n+3.\dim S(n)/(\text{in}_{>}I_{G_{n}^{t}})=n+3.

To see that the dimension is at least n+3n+3, we construct a chain of prime ideals in S(n)S(n) containing in>IGnt\text{in}_{>}I_{G_{n}^{t}}. Since every monomial generator of in>IGnt\text{in}_{>}I_{G_{n}^{t}} contains a variable of odd index, we begin with Pn=({xkk odd, 2<k<2n+4})P_{n}=(\{x_{k}\mid\text{$k$ odd, $2<k<2n+4$}\}), a prime ideal containing in>IGnt\text{in}_{>}I_{G_{n}^{t}}. Then we have the chain of prime ideals PnPn+(x0)Pn+(x0,x2)Pn+(x0,x2,x4)Pn+({x2i0in+2})P_{n}\subsetneq P_{n}+(x_{0})\subsetneq P_{n}+(x_{0},x_{2})\subsetneq P_{n}+(x_{0},x_{2},x_{4})\subsetneq\cdots\subsetneq P_{n}+(\{x_{2i}\mid 0\leq i\leq n+2\}), so that

dimS(n)/(in>IGnt)n+3.\dim S(n)/(\text{in}_{>}I_{G_{n}^{t}})\geq n+3.

To see that the dimension is at most n+3n+3, we find a sequence of n+3n+3 elements in S(n)/(in>IGnt)S(n)/(\text{in}_{>}I_{G_{n}^{t}}) such that the quotient by the ideal they generate has dimension zero. Let

Xn=x0,x2x3,x4x5,,x2nx2n+1,x2n+2x2n+3,x2n+4X_{n}=x_{0},x_{2}-x_{3},x_{4}-x_{5},\ldots,x_{2n}-x_{2n+1},x_{2n+2}-x_{2n+3},x_{2n+4}

in S(n)S(n), and take the quotient of S(n)/(in>IGnt)S(n)/(\text{in}_{>}I_{G_{n}^{t}}) by the image of XnX_{n} to obtain the following. In the last step, we rewrite the quotient of S(n)S(n) and (in>IGnt)+(Xn)(\text{in}_{>}I_{G_{n}^{t}})+(X_{n}) by (Xn)(X_{n}) by setting x0x_{0} and x2n+4x_{2n+4} equal to zero and replacing x2ix_{2i} with x2i+1x_{2i+1} for 1in+11\leq i\leq n+1:

S(n)/(in>IGnt)((in>IGnt)+(Xn))/(in>IGnt)\displaystyle\frac{S(n)/(\text{in}_{>}I_{G_{n}^{t}})}{((\text{in}_{>}I_{G_{n}^{t}})+(X_{n}))/(\text{in}_{>}I_{G_{n}^{t}})} \displaystyle\cong S(n)((in>IGnt)+(Xn))\displaystyle\frac{S(n)}{((\text{in}_{>}I_{G_{n}^{t}})+(X_{n}))}
\displaystyle\cong S(n)/(Xn)((in>IGnt)+(Xn))/(Xn)\displaystyle\frac{S(n)/(X_{n})}{((\text{in}_{>}I_{G_{n}^{t}})+(X_{n}))/(X_{n})}
\displaystyle\cong k[x3,x5,,x2n+1,x2n+3](x32,{x2i+1x2i+3,x2i+321in}).\displaystyle\frac{k[x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}]}{(x_{3}^{2},\{x_{2i+1}x_{2i+3},x_{2i+3}^{2}\mid 1\leq i\leq n\})}.

Since

(x32,{x2i+1x2i+3,x2i+321in})=(x3,x5,,x2n+3),\sqrt{(x_{3}^{2},\{x_{2i+1}x_{2i+3},x_{2i+3}^{2}\mid 1\leq i\leq n\})}=(x_{3},x_{5},\ldots,x_{2n+3}),

the above ring has dimension zero. Thus,

dimS(n)/(in>IGnt)n+3.\dim S(n)/(\text{in}_{>}I_{G_{n}^{t}})\leq n+3.

We conclude that dimR(n,t)=dimS(n)/(in>IGnt)=n+3\dim R(n,t)=\dim S(n)/(\text{in}_{>}I_{G_{n}^{t}})=n+3. ∎

Corollary 3.5.

The projective dimension of R(n,t)R(n,t) over S(n)S(n) is

pdS(n)R(n,t)=n+1.\textnormal{pd}_{S(n)}R(n,t)=n+1.
Proof.

We know the Krull dimension of the polynomial ring S(n)S(n) is 2n+42n+4. The result follows from the fact that R(n,t)R(n,t) is Cohen-Macaulay (Corollary 2.16) and from the graded version of the Auslander-Buchsbaum formula. ∎

Remark 3.6.

The proof of the previous theorem shows that the image of

Xn=x0,x2x3,x4x5,,x2nx2n+1,x2n+2x2n+3,x2n+4X_{n}=x_{0},x_{2}-x_{3},x_{4}-x_{5},\ldots,x_{2n}-x_{2n+1},x_{2n+2}-x_{2n+3},x_{2n+4}

in S(n)/(in>IGnt)S(n)/(\text{in}_{>}I_{G_{n}^{t}}) is a system of parameters for S(n)/(in>IGnt)S(n)/(\text{in}_{>}I_{G_{n}^{t}}). We prove in the next theorem that the image of XnX_{n} in R(n,t)R(n,t) (which we call Xn¯\overline{X_{n}}) is also a system of parameters for R(n,t)R(n,t). Before doing so, we introduce some notation and a definition which allows us to better grapple with the quotient ring R(n,t)/(Xn¯)R(n,t)/(\overline{X_{n}}).

Definition 3.7.

Consider the isomorphism

R(n,t)(Xn¯)=S(n)/(IGnt)((IGnt)+(Xn))/(IGnt)S(n)/(Xn)((IGnt)+(Xn))/(Xn)\frac{R(n,t)}{(\overline{X_{n}})}=\frac{S(n)/(I_{G_{n}^{t}})}{((I_{G_{n}^{t}})+(X_{n}))/(I_{G_{n}^{t}})}\cong\frac{S(n)/(X_{n})}{((I_{G_{n}^{t}})+(X_{n}))/(X_{n})}

We view taking the quotient by XnX_{n} as setting x0x_{0} and x2n+4x_{2n+4} equal to zero and replacing x2ix_{2i} with x2i+1x_{2i+1} for 1in+11\leq i\leq n+1 to obtain

S(n)^:=k[x3,x5,,x2n+1,x2n+3]S(n)/(Xn).\widehat{S(n)}:=k[x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}]\cong S(n)/(X_{n}).

By the same process (detailed below), we obtain the ideal IGnt^(IGnt+(Xn))/(Xn)\widehat{I_{G_{n}^{t}}}\cong(I_{G_{n}^{t}}+(X_{n}))/(X_{n}). We further define the quotient

R(n,t)^:=S(n)^/IGnt^R(n,t)/(Xn¯).\widehat{R(n,t)}:=\widehat{S(n)}/\widehat{I_{G_{n}^{t}}}\cong R(n,t)/(\overline{X_{n}}).

We find this notation natural since it is often used for the removal of variables, and the quotient by XnX_{n} may be viewed as identifying and removing variables. Since this work has no completions in it, there should be no conflict of notation.

Definition 3.8.

To define IGnt^\widehat{I_{G_{n}^{t}}} in particular, we recall the standard generators of IGntI_{G_{n}^{t}} and introduce further notation to describe the generators of IGnt^(IGnt+(Xn))/(Xn)\widehat{I_{G_{n}^{t}}}\cong(I_{G_{n}^{t}}+(X_{n}))/(X_{n}). By Remark 2.17, the standard generators of IGntI_{G_{n}^{t}} are

s1\displaystyle s_{1} =\displaystyle= x2x3xj1x4\displaystyle x_{2}x_{3}-x_{j_{1}}x_{4}
s2i\displaystyle s_{2i} =\displaystyle= x2i+1x2i+3xj2ix2i+4\displaystyle x_{2i+1}x_{2i+3}-x_{j_{2i}}x_{2i+4}
s2i+1\displaystyle s_{2i+1} =\displaystyle= x2i+2x2i+3xj2i+1x2i+4,\displaystyle x_{2i+2}x_{2i+3}-x_{j_{2i+1}}x_{2i+4},

for 1in1\leq i\leq n, where the nonnegative integers jkj_{k} are as in Remark 2.11.

Let ι^\widehat{\iota} be the largest index such that j2ι^=0j_{2\widehat{\iota}}=0. By Remark 2.11, we see that the j2ij_{2i} are defined recursively and form a non-decreasing sequence. Then

j1=j2=j4=j6==j2ι^=0,j_{1}=j_{2}=j_{4}=j_{6}=\cdots=j_{2\widehat{\iota}}=0,

and since we view taking the quotient by XnX_{n} as setting x0x_{0} and x2n+4x_{2n+4} equal to zero and replacing x2ix_{2i} with x2i+1x_{2i+1} for 1in+11\leq i\leq n+1, we define IGnt^\widehat{I_{G_{n}^{t}}} by replacing xjkx_{j_{k}} with xJkx_{J_{k}} (defined below) for 1k<2n1\leq k<2n to obtain

s1^\displaystyle\widehat{s_{1}} =\displaystyle= x32xJ1x5\displaystyle x_{3}^{2}-x_{J_{1}}x_{5}
s2i^\displaystyle\widehat{s_{2i}} =\displaystyle= x2i+1x2i+3xJ2ix2i+5\displaystyle x_{2i+1}x_{2i+3}-x_{J_{2i}}x_{2i+5}
s2i+1^\displaystyle\widehat{s_{2i+1}} =\displaystyle= x2i+32xJ2i+1x2i+5\displaystyle x_{2i+3}^{2}-x_{J_{2i+1}}x_{2i+5}
s2n^\displaystyle\widehat{s_{2n}} =\displaystyle= x2n+1x2n+3\displaystyle x_{2n+1}x_{2n+3}
s2n+1^\displaystyle\widehat{s_{2n+1}} =\displaystyle= x2n+32\displaystyle x_{2n+3}^{2}

for 1i<n1\leq i<n, where

xJk={0 if k is even and k2ι^, or if k=1xjk+1 if 2ι^<k<2n and jk is evenxjk if jk is oddx_{J_{k}}=\begin{cases}0&\text{ if }k\text{ is even and }k\leq 2\widehat{\iota},\text{ or if }k=1\\ x_{j_{k}+1}&\text{ if }2\widehat{\iota}<k<2n\text{ and $j_{k}$ is even}\\ x_{j_{k}}&\text{ if }j_{k}\text{ is odd}\end{cases}

We note that JkkJ_{k}\leq k for each 1k<2n1\leq k<2n, since jkk1j_{k}\leq k-1 by Remark 2.11. By properties of the original jkj_{k} from Remark 2.11, we know that xJ1=xJ2=0x_{J_{1}}=x_{J_{2}}=0, J3=3J_{3}=3, and for 2i<n2\leq i<n,

ti+1\displaystyle t_{i+1} =\displaystyle= 0xJ2i=xJ2i2J2i+1=2i1\displaystyle 0\iff x_{J_{2i}}=x_{J_{2i-2}}\iff J_{2i+1}=2i-1
ti+1\displaystyle t_{i+1} =\displaystyle= 1xJ2i=xJ2i1J2i+1=2i+1.\displaystyle 1\iff x_{J_{2i}}=x_{J_{2i-1}}\iff J_{2i+1}=2i+1.
Example 3.9.

We construct the ring R(2,(0,0,0))R(2,(0,0,0)). For the graph G2(0,0,0)G_{2}^{(0,0,0)}\in\mathcal{F}, we have the toric ring

R(2,(0,0,0))=k[x0,x2,x3,,x8](x2x3x0x4,x3x5x0x6,x4x5x2x6,x5x7x0x8,x6x7x3x8)R(2,(0,0,0))=\frac{k[x_{0},x_{2},x_{3},\ldots,x_{8}]}{(x_{2}x_{3}-x_{0}x_{4},x_{3}x_{5}-x_{0}x_{6},x_{4}x_{5}-x_{2}x_{6},x_{5}x_{7}-x_{0}x_{8},x_{6}x_{7}-x_{3}x_{8})}

coming from the ladder-like structure

L2(0,0,0)=x0x2x5x3x4x6x7x8L_{2}^{(0,0,0)}=\begin{matrix}x_{0}&x_{2}&x_{5}\\ x_{3}&x_{4}&x_{6}\\ x_{7}&&x_{8}\end{matrix}

from Example 2.3. We know

X2=x0,x2x3,x4x5,,x4x5,x6x7,x8,X_{2}=x_{0},x_{2}-x_{3},x_{4}-x_{5},\ldots,x_{4}-x_{5},x_{6}-x_{7},x_{8},

so that R(2,(0,0,0))/(X2¯)R(2,(0,0,0))/(\overline{X_{2}}) is isomorphic to

k[x0,x2,x3,x4,x5,x6,x7,x8](x2x3x0x4,x3x5x0x6,x4x5x2x6,x5x7x0x8,x6x7x3x8,x0,x2x3,,x8)\frac{k[x_{0},x_{2},x_{3},x_{4},x_{5},x_{6},x_{7},x_{8}]}{(x_{2}x_{3}-x_{0}x_{4},x_{3}x_{5}-x_{0}x_{6},x_{4}x_{5}-x_{2}x_{6},x_{5}x_{7}-x_{0}x_{8},x_{6}x_{7}-x_{3}x_{8},x_{0},x_{2}-x_{3},\ldots,x_{8})}
k[x3,x5,x7](x32,x3x5,x52x3x7,x5x7,x72)=R(2,(0,0,0))^.\cong\frac{k[x_{3},x_{5},x_{7}]}{(x_{3}^{2},x_{3}x_{5},x_{5}^{2}-x_{3}x_{7},x_{5}x_{7},x_{7}^{2})}=\widehat{R(2,(0,0,0))}.

Now we show that XnX_{n} is also a system of parameters for R(n,t)R(n,t), and not just for the quotient by the initial ideal.

Proposition 3.10.

Let R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}} and let

Xn=x0,x2x3,x4x5,,x2nx2n+1,x2n+2x2n+3,x2n+4X_{n}=x_{0},x_{2}-x_{3},x_{4}-x_{5},\ldots,x_{2n}-x_{2n+1},x_{2n+2}-x_{2n+3},x_{2n+4}

so that the image of XnX_{n} in S(n)/(in>IGnt)S(n)/(\text{in}_{>}I_{G_{n}^{t}}) is the system of parameters from Remark 3.6. Then the image of XnX_{n} in R(n,t)R(n,t) is a system of parameters for R(n,t)R(n,t).

Proof.

Let XnX_{n} be defined as above. Then by Theorem 3.4 and Definition 3.7 we need only show that dimR(n,t)^=0\dim\widehat{R(n,t)}=0. We have for n=0n=0 that

R(0,t)^=k[x3](x32),\widehat{R(0,t)}=\frac{k[x_{3}]}{(x_{3}^{2})},

for n=1n=1

R(1,t)^=k[x3,x5](x32,x3x5,x52),\widehat{R(1,t)}=\frac{k[x_{3},x_{5}]}{(x_{3}^{2},x_{3}x_{5},x_{5}^{2})},

and for n>1n>1

R(n,t)^=S(n)^IGnt^=k[x3,x5,,x2n+1,x2n+3]({s1^,s2i^,s2i+1^1in}),\widehat{R(n,t)}=\frac{\widehat{S(n)}}{\widehat{I_{G_{n}^{t}}}}=\frac{k[x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}]}{(\{\widehat{s_{1}},\widehat{s_{2i}},\widehat{s_{2i+1}}\mid 1\leq i\leq n\})},

where

s1^\displaystyle\widehat{s_{1}} =\displaystyle= x32\displaystyle x_{3}^{2}
s2i^\displaystyle\widehat{s_{2i}} =\displaystyle= x2i+1x2i+3xJ2ix2i+5\displaystyle x_{2i+1}x_{2i+3}-x_{J_{2i}}x_{2i+5}
s2i+1^\displaystyle\widehat{s_{2i+1}} =\displaystyle= x2i+32xJ2i+1x2i+5\displaystyle x_{2i+3}^{2}-x_{J_{2i+1}}x_{2i+5}
s2n^\displaystyle\widehat{s_{2n}} =\displaystyle= x2n+1x2n+3\displaystyle x_{2n+1}x_{2n+3}
s2n+1^\displaystyle\widehat{s_{2n+1}} =\displaystyle= x2n+32\displaystyle x_{2n+3}^{2}

for 1i<n1\leq i<n from Definition 3.8. We know dimR(n,t)^=dimS(n)^/IGnt^=dimS(n)^/IGnt^\dim\widehat{R(n,t)}=\dim\widehat{S(n)}/\widehat{I_{G_{n}^{t}}}=\dim\widehat{S(n)}/\sqrt{\widehat{I_{G_{n}^{t}}}}. We claim that

IGnt^=(x3,x5,,x2n+1,x2n+3).\sqrt{\widehat{I_{G_{n}^{t}}}}=\left(x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}\right).

This is clear for n{0,1}n\in\{0,1\}. For n>1n>1, we prove this by induction. Since s1^=x32\widehat{s_{1}}=x_{3}^{2} and s2n+1^=x2n+32\widehat{s_{2n+1}}=x_{2n+3}^{2} are in IGnt^\widehat{I_{G_{n}^{t}}}, we have x3,x2n+3IGnt^x_{3},x_{2n+3}\in\sqrt{\widehat{I_{G_{n}^{t}}}}. Since

s3^=x52x3x7IGnt^IGnt^\widehat{s_{3}}=x_{5}^{2}-x_{3}x_{7}\in\widehat{I_{G_{n}^{t}}}\subseteq\sqrt{\widehat{I_{G_{n}^{t}}}}

and x3IGnt^x_{3}\in\sqrt{\widehat{I_{G_{n}^{t}}}}, we get x52IGnt^x_{5}^{2}\in\sqrt{\widehat{I_{G_{n}^{t}}}}, so that x5IGnt^x_{5}\in\sqrt{\widehat{I_{G_{n}^{t}}}}. Now suppose x2i1,x2i+1IGnt^x_{2i-1},x_{2i+1}\in\sqrt{\widehat{I_{G_{n}^{t}}}} for 2i<n2\leq i<n. We have

s2i+1^=x2i+32xJ2i+1x2i+5IGnt^IGnt^.\widehat{s_{2i+1}}=x_{2i+3}^{2}-x_{J_{2i+1}}x_{2i+5}\in\widehat{I_{G_{n}^{t}}}\subseteq\sqrt{\widehat{I_{G_{n}^{t}}}}.

But xJ2i+1{x2i1,x2i+1}x_{J_{2i+1}}\in\{x_{2i-1},x_{2i+1}\} by Definition 3.8 and {x2i1,x2i+1}IGnt^\{x_{2i-1},x_{2i+1}\}\subseteq\sqrt{\widehat{I_{G_{n}^{t}}}} by induction, so that x2i+32IGnt^x_{2i+3}^{2}\in\sqrt{\widehat{I_{G_{n}^{t}}}}, and hence x2i+3IGnt^x_{2i+3}\in\sqrt{\widehat{I_{G_{n}^{t}}}}. We conclude that

(x3,x5,,x2n+1,x2n+3)IGnt^(x3,x5,,x2n+1,x2n+3),\left(x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}\right)\subseteq\sqrt{\widehat{I_{G_{n}^{t}}}}\subseteq\left(x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}\right),

so we have equality. Since S(n)^/IGnt^k\widehat{S(n)}/\sqrt{\widehat{I_{G_{n}^{t}}}}\cong k has dimension zero, so does R(n,t)^R(n,t)/(Xn¯)\widehat{R(n,t)}\cong R(n,t)/(\overline{X_{n}}). Thus, the image of XnX_{n} is a system of parameters for R(n,t)R(n,t). ∎

Remark 3.11.

We note that as a consequence of the proof of the preceding theorem, the ring R(n,t)^\widehat{R(n,t)} is Artinian, which will be relevant in Section 3.2.

Corollary 3.12.

The image of

Xn=x0,x2x3,x4x5,,x2nx2n+1,x2n+2x2n+3,x2n+4X_{n}=x_{0},x_{2}-x_{3},x_{4}-x_{5},\ldots,x_{2n}-x_{2n+1},x_{2n+2}-x_{2n+3},x_{2n+4}

in R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}} is a regular sequence for R(n,t)R(n,t).

Proof.

We know by Proposition 3.10 that the image of XnX_{n} in R(n,t)R(n,t) is a linear system of parameters. Since the rings R(n,t)R(n,t) are Cohen-Macaulay (Corollary 2.16), we are done. ∎

3.2. Length, Multiplicity, and Regularity

In this section, we determine the multiplicity and Castelnuovo-Mumford regularity of the toric rings R(n,t)R(n,t) coming from the associated graphs GntG_{n}^{t}\in\mathcal{F} by computing the length of the Artinian rings

R(n,t)^R(n,t)/(Xn¯)\widehat{R(n,t)}\cong R(n,t)/(\overline{X_{n}})

from Definition 3.7. We know by Corollary 3.12 that Xn¯\overline{X_{n}} is a linear regular sequence for R(n,t)R(n,t), which allows us to compute the multiplicity of the rings R(n,t)R(n,t). As a corollary of Theorem 3.16, which establishes the Hilbert function for R(n,t)^\widehat{R(n,t)}, we obtain the multiplicity and regularity of R(n,t)R(n,t). We also develop an alternate graph-theoretic proof for the regularity of R(n,t)R(n,t), which is included at the end of this section.

We begin with a lemma establishing a vector space basis for R(n,t)^\widehat{R(n,t)}, which we use extensively for our results.

Lemma 3.13.

The image of all squarefree monomials with only odd indices whose indices are at least four apart, together with the image of 1k1_{k}, forms a vector space basis for R(n,t)^\widehat{R(n,t)} over kk.

Proof.

We recall for the reader the definition of R(n,t)^\widehat{R(n,t)} and then find the initial ideal of IGnt^\widehat{I_{G_{n}^{t}}} and use Macaulay’s Basis Theorem to show that the desired representatives form a basis for R(n,t)^\widehat{R(n,t)} as a vector space over kk.

From Definition 3.7, we have

R(n,t)/(Xn¯)R(n,t)^=S(n)^/IGnt^,R(n,t)/(\overline{X_{n}})\cong\widehat{R(n,t)}=\widehat{S(n)}/\widehat{I_{G_{n}^{t}}},

where

S(n)^=k[x3,x5,,x2n+1,x2n+3].\widehat{S(n)}=k[x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}].

By Definition 3.8, for 1i<n1\leq i<n the ideal IGnt^\widehat{I_{G_{n}^{t}}} is generated by

s1^\displaystyle\widehat{s_{1}} =\displaystyle= x32xJ1x5\displaystyle x_{3}^{2}-x_{J_{1}}x_{5}
s2i^\displaystyle\widehat{s_{2i}} =\displaystyle= x2i+1x2i+3xJ2ix2i+5\displaystyle x_{2i+1}x_{2i+3}-x_{J_{2i}}x_{2i+5}
s2i+1^\displaystyle\widehat{s_{2i+1}} =\displaystyle= x2i+32xJ2i+1x2i+5\displaystyle x_{2i+3}^{2}-x_{J_{2i+1}}x_{2i+5}
s2n^\displaystyle\widehat{s_{2n}} =\displaystyle= x2n+1x2n+3\displaystyle x_{2n+1}x_{2n+3}
s2n+1^\displaystyle\widehat{s_{2n+1}} =\displaystyle= x2n+32,\displaystyle x_{2n+3}^{2},

where xJ1=xJ2=0x_{J_{1}}=x_{J_{2}}=0, J3=3J_{3}=3, and for 2i<n2\leq i<n,

ti+1\displaystyle t_{i+1} =\displaystyle= 0xJ2i=xJ2i2J2i+1=2i1\displaystyle 0\iff x_{J_{2i}}=x_{J_{2i-2}}\iff J_{2i+1}=2i-1
ti+1\displaystyle t_{i+1} =\displaystyle= 1xJ2i=xJ2i1J2i+1=2i+1.\displaystyle 1\iff x_{J_{2i}}=x_{J_{2i-1}}\iff J_{2i+1}=2i+1.

We first show that the image of the monomials with the desired property is a basis in the quotient of S(n)^\widehat{S(n)} by the initial ideal in>IGnt^\text{in}_{>}\widehat{I_{G_{n}^{t}}}. By Macaulay’s Basis Theorem, which is Theorem 1.5.7 in [21], the image of these monomials in R(n,t)^\widehat{R(n,t)} is also a basis.

To find the initial ideal of IGnt^\widehat{I_{G_{n}^{t}}}, we establish that the given generators sk^\widehat{s_{k}} are a Gröbner basis for IGnt^\widehat{I_{G_{n}^{t}}} with respect to the degree reverse lexicographic order >>. This is a relatively straightforward computation using Buchberger’s Criterion, with separate cases for when xJk=0x_{J_{k}}=0.

If we adopt SS-polynomial notation Si,jS_{i,j} for the SS-polynomial of si^\widehat{s_{i}} and sj^\widehat{s_{j}}, then the cases to consider are

S1,2,S2,3,S2,4,S2n2,2n,S2n1,2n,S2n,2n+1S_{1,2},S_{2,3},S_{2,4},S_{2n-2,2n},S_{2n-1,2n},S_{2n,2n+1}
S2i1,2i for 1<i<nS_{2i-1,2i}\text{ for }1<i<n
S2i,2i+1 for 1<i<nS_{2i,2i+1}\text{ for }1<i<n
S2i,2i+2 for 1<i<n1.S_{2i,2i+2}\text{ for }1<i<n-1.

To give a flavor of the computation involved, we show the case S2i,2i+2S_{2i,2i+2} for 1<i<n11<i<n-1, and leave the remaining cases to the reader. We show in each subcase that S2i,2i+2S_{2i,2i+2} is equal to zero or to a sum of basis elements with coefficients in S(n)^\widehat{S(n)}, so that the reduced form of S2i,2i+2S_{2i,2i+2} is zero in each subcase. We have

S2i,2i+2\displaystyle S_{2i,2i+2} =\displaystyle= x2i+5(x2i+1x2i+3xJ2ix2i+5)x2i+1(x2i+3x2i+5xJ2i+2x2i+7)\displaystyle x_{2i+5}(x_{2i+1}x_{2i+3}-x_{J_{2i}}x_{2i+5})-x_{2i+1}(x_{2i+3}x_{2i+5}-x_{J_{2i+2}}x_{2i+7})
=\displaystyle= xJ2ix2i+52+x2i+1xJ2i+2x2i+7\displaystyle-x_{J_{2i}}x_{2i+5}^{2}+x_{2i+1}x_{J_{2i+2}}x_{2i+7}

Case 1: If ti+2=0t_{i+2}=0, then xJ2i+2=xJ2ix_{J_{2i+2}}=x_{J_{2i}} and J2i+3=2i+1J_{2i+3}=2i+1, so we have

S2i,2i+2\displaystyle S_{2i,2i+2} =\displaystyle= xJ2ix2i+52+xJ2ix2i+1x2i+7\displaystyle-x_{J_{2i}}x_{2i+5}^{2}+x_{J_{2i}}x_{2i+1}x_{2i+7}
=\displaystyle= xJ2is2i+3^\displaystyle-x_{J_{2i}}\widehat{s_{2i+3}}

We note that if xJ2i=xJ2i+2=0x_{J_{2i}}=x_{J_{2i+2}}=0, then S2i,2i+2=0S_{2i,2i+2}=0.

Case 2: If ti+2=1t_{i+2}=1, then xJ2i+2=xJ2i+1x_{J_{2i+2}}=x_{J_{2i+1}} and J2i+3=2i+3J_{2i+3}=2i+3, so we have

S2i,2i+2\displaystyle S_{2i,2i+2} =\displaystyle= xJ2ix2i+52+x2i+1xJ2i+1x2i+7.\displaystyle-x_{J_{2i}}x_{2i+5}^{2}+x_{2i+1}x_{J_{2i+1}}x_{2i+7}.

Case 2.1: If in addition ti+1=0t_{i+1}=0, then xJ2i=xJ2i2x_{J_{2i}}=x_{J_{2i-2}} and J2i+1=2i1J_{2i+1}=2i-1, so we have

S2i,2i+2\displaystyle S_{2i,2i+2} =\displaystyle= xJ2i2x2i+52+x2i+1x2i1x2i+7\displaystyle-x_{J_{2i-2}}x_{2i+5}^{2}+x_{2i+1}x_{2i-1}x_{2i+7}
=\displaystyle= xJ2i2(x2i+52xJ2i+3x2i+7)+x2i+7(x2i1x2i+1xJ2i2x2i+3)\displaystyle-x_{J_{2i-2}}(x_{2i+5}^{2}-x_{J_{2i+3}}x_{2i+7})+x_{2i+7}(x_{2i-1}x_{2i+1}-x_{J_{2i-2}}x_{2i+3})
=\displaystyle= xJ2i2s2i+3^+x2i+7s2i2^.\displaystyle-x_{J_{2i-2}}\widehat{s_{2i+3}}+x_{2i+7}\widehat{s_{2i-2}}.

We note that if xJ2i=xJ2i2=0x_{J_{2i}}=x_{J_{2i-2}}=0, then S2i,2i+2=x2i+7s2i2^S_{2i,2i+2}=x_{2i+7}\widehat{s_{2i-2}}.

Case 2.2: If in addition ti+1=1t_{i+1}=1, then xJ2i=xJ2i1x_{J_{2i}}=x_{J_{2i-1}} and J2i+1=2i+1J_{2i+1}=2i+1, so we have

S2i,2i+2\displaystyle S_{2i,2i+2} =\displaystyle= xJ2i1x2i+52+x2i+12x2i+7\displaystyle-x_{J_{2i-1}}x_{2i+5}^{2}+x_{2i+1}^{2}x_{2i+7}
=\displaystyle= xJ2i1(x2i+52xJ2i+3x2i+7)+x2i+7(x2i+12xJ2i1x2i+3)\displaystyle-x_{J_{2i-1}}(x_{2i+5}^{2}-x_{J_{2i+3}}x_{2i+7})+x_{2i+7}(x_{2i+1}^{2}-x_{J_{2i-1}}x_{2i+3})
=\displaystyle= xJ2i1s2i+3^+x2i+7s2i1^.\displaystyle-x_{J_{2i-1}}\widehat{s_{2i+3}}+x_{2i+7}\widehat{s_{2i-1}}.

This concludes the case S2i,2i+2S_{2i,2i+2} for 1<i<n11<i<n-1. The remaining cases are similar.

Then the given generators sk^\widehat{s_{k}} are a Gröbner Basis for IGnt^\widehat{I_{G_{n}^{t}}}, so that the initial ideal is

in>(IGnt^)=(x32,{x2i+1x2i+3,x2i+321in})\text{in}_{>}(\widehat{I_{G_{n}^{t}}})=(x_{3}^{2},\{x_{2i+1}x_{2i+3},x_{2i+3}^{2}\mid 1\leq i\leq n\})

in the ring S(n)^=k[x3,x5,,x2n+1,x2n+3]\widehat{S(n)}=k[x_{3},x_{5},\ldots,x_{2n+1},x_{2n+3}]. Since in>(IGnt^)\text{in}_{>}(\widehat{I_{G_{n}^{t}}}) consists precisely of all squares of variables in S(n)^\widehat{S(n)} and all degree two products of variables whose indices differ by exactly two, it follows that the image of the squarefree monomials whose indices are at least four apart, together with the image of 1k1_{k}, forms a basis for S(n)^in>IGnt^\frac{\widehat{S(n)}}{\text{in}_{>}\widehat{I_{G_{n}^{t}}}}. By Macaulay’s Basis Theorem, the image of these monomials in R(n,t)^=S(n)^IGnt^\widehat{R(n,t)}=\frac{\widehat{S(n)}}{\widehat{I_{G_{n}^{t}}}} is also a basis.

We use the lemma above to establish facts about the vector space dimensions of degree ii pieces of R(n,t)^\widehat{R(n,t)}, which are applied further below to establish length and multiplicity.

Notation 3.14.

Throughout this section, we use dn,i:=dimk(R(n,t)^)id_{n,i}:=\dim_{k}(\widehat{R(n,t)})_{i} for the vector space dimension of the degree ii piece of R(n,t)^\widehat{R(n,t)}, that is, for the iith coefficient in the Hilbert series of R(n,t)^\widehat{R(n,t)}. By Lemma 3.13, these are independent of tt.

We establish a recursive relationship between these dimensions by introducing a short exact sequence of vector spaces.

Lemma 3.15.

For n2n\geq 2 and i1i\geq 1, the vector space dimension dn,i=dimk(R(n,t)^)id_{n,i}=\dim_{k}(\widehat{R(n,t)})_{i} satisfies the recursive relationship

dn,i=dn1,i+dn2,i1.d_{n,i}=d_{n-1,i}+d_{n-2,i-1}.
Proof.

We use the vector space basis defined in Lemma 3.13. We note that the basis elements described are actually monomial representatives (which do not depend on tt) of equivalence classes (which do depend on tt), but we suppress this and speak as if they are monomials, not depending on tt. We then take the liberty of suppressing tt in what follows, for convenience. We recall for the reader that

S(n)^\displaystyle\widehat{S(n)} =\displaystyle= k[x3,x5,,x2n3,x2n1,x2n+1,x2n+3]\displaystyle k[x_{3},x_{5},\ldots,x_{2n-3},x_{2n-1},x_{2n+1},x_{2n+3}]
S(n1)^\displaystyle\widehat{S(n-1)} =\displaystyle= k[x3,x5,,x2n3,x2n1,x2n+1]\displaystyle k[x_{3},x_{5},\ldots,x_{2n-3},x_{2n-1},x_{2n+1}]
S(n2)^\displaystyle\widehat{S(n-2)} =\displaystyle= k[x3,x5,,x2n3,x2n1]\displaystyle k[x_{3},x_{5},\ldots,x_{2n-3},x_{2n-1}]

Let x2n+3:(R(n2)^)i1(R(n)^)ix_{2n+3}:(\widehat{R(n-2)})_{i-1}\to(\widehat{R(n)})_{i} be multiplication by x2n+3x_{2n+3}, and let
x2n+3^:(R(n)^)i(R(n1)^)i\widehat{x_{2n+3}}:(\widehat{R(n)})_{i}\to(\widehat{R(n-1)})_{i} be defined for a basis element bb by

x2n+3^(b)={bif x2n+3b0if x2n+3b.\widehat{x_{2n+3}}(b)=\begin{cases}b&\text{if }x_{2n+3}\nmid b\\ 0&\text{if }x_{2n+3}\mid b.\end{cases}

We note that these vector space maps are well-defined, since 1k1_{k} or a squarefree monomial with odd indices at least four apart has an output of 0, 11, or a monomial with the same properties. The following sequence of vector spaces is exact

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(R(n2)^)i1\textstyle{(\widehat{R(n-2)})_{i-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}x2n+3\scriptstyle{x_{2n+3}}(R(n)^)i\textstyle{(\widehat{R(n)})_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}x2n+3^\scriptstyle{\widehat{x_{2n+3}}}(R(n1)^)i\textstyle{(\widehat{R(n-1)})_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,}

so that

dn,i=dn1,i+dn2,i1.d_{n,i}=d_{n-1,i}+d_{n-2,i-1}.

Applying Lemma 3.15 and induction, we achieve the following closed formula for the coefficients of the Hilbert series of R(n,t)^\widehat{R(n,t)}.

Theorem 3.16.

If R(n,t)=S(n)/IGntR(n,t)=S(n)/I_{G_{n}^{t}} and R(n,t)^R(n,t)/(Xn¯)\widehat{R(n,t)}\cong R(n,t)/(\overline{X_{n}}), we have

dimk(R(n,t)^)i={1i=01i!j=1i(n+j2(i1))i1.{\displaystyle\dim_{k}(\widehat{R(n,t)})_{i}=\begin{cases}1&i=0\\ {\displaystyle\frac{1}{i!}\prod_{j=1}^{i}(n+j-2(i-1))}&i\geq 1.\end{cases}}

In particular, dimk(R(n,t)^)i=0\dim_{k}(\widehat{R(n,t)})_{i}=0 when i>n/2+1i>n/2+1.

Proof.

We begin with the proof of the last statement, and note that throughout the following proof we use Notation 3.14. When n0n\geq 0 is even, i=n/2+2i=n/2+2, and j=2(i)j=2(\leq i), we have a factor of zero, and when n1n\geq 1 is odd, i=n/2+3/2i=n/2+3/2, and j=1(<i)j=1(<i), we have a factor of zero. Thus dn,i=0d_{n,i}=0 when i>n/2+1i>n/2+1.

Now we establish the base cases i,n{0,1}i,n\in\{0,1\}, then proceed by induction. It is clear that dn,0=1d_{n,0}=1, generated by 1k1_{k}. By Lemma 3.13 and by the fact that R(n,t)^\widehat{R(n,t)} is a graded quotient, every nonzero element of positive degree ii can be represented uniquely as a sum of degree ii squarefree monomials with odd indices whose indices are at least four apart. Then (R(n,t)^)1(\widehat{R(n,t)})_{1} is generated by the images of all the odd variables

x3,x2(1)+3,,x2n+3x_{3},x_{2(1)+3},\ldots,x_{2n+3}

in S(n)S(n), so that

dn,1=n+1=11!j=11(n+j2(11))d_{n,1}=n+1=\frac{1}{1!}\prod_{j=1}^{1}(n+j-2(1-1))

matches the given formula.

Now we establish the base cases n=0n=0 and n=1n=1 for all ii. We recognize that the first monomial of degree two with odd indices at least four apart is x3x7x_{3}x_{7}, which does not exist until n=2n=2, so we have

d0,i={1i=01i=10else={1i=01i!j=1i(j2(i1))i1d_{0,i}=\begin{cases}1&i=0\\ 1&i=1\\ 0&\text{else}\end{cases}=\begin{cases}1&i=0\\ {\displaystyle\frac{1}{i!}\prod_{j=1}^{i}(j-2(i-1))}&i\geq 1\end{cases}

and

d1,i={1i=02i=10else={1i=01i!j=1i(1+j2(i1))i1,d_{1,i}=\begin{cases}1&i=0\\ 2&i=1\\ 0&\text{else}\end{cases}=\begin{cases}1&i=0\\ {\displaystyle\frac{1}{i!}\prod_{j=1}^{i}(1+j-2(i-1))}&i\geq 1,\end{cases}

which match the given formula.

This gives us the following table of base cases for dn,id_{n,i}, which match the given formula:

nin\setminus i 0 1 2 3 4 5 6 7 \cdots
0 1 1 0 0 0 0 0 0 \cdots
1 1 2 0 0 0 0 0 0 \cdots
2 1 3
3 1 4
4 1 5
\vdots \vdots \vdots

We recall by Lemma 3.15 that we have the recursive relationship

dn,i=dn1,i+dn2,i1d_{n,i}=d_{n-1,i}+d_{n-2,i-1}

for n2n\geq 2 and i1i\geq 1. We proceed by induction. Suppose N,I2N,I\geq 2 and that the dimension formula holds for all ii when n<Nn<N. By our recursion and by induction, we have

dN,I\displaystyle d_{N,I} =\displaystyle= dN1,I+dN2,I1\displaystyle d_{N-1,I}+d_{N-2,I-1}
=\displaystyle= 1I!j=1I(N1+j2(I1))+1(I1)!j=1I1(N2+j2(I2))\displaystyle\frac{1}{I!}\prod_{j=1}^{I}(N-1+j-2(I-1))+\frac{1}{(I-1)!}\prod_{j=1}^{I-1}(N-2+j-2(I-2))
=\displaystyle= 1I!(N2(I1))j=1I1(N+j2(I1))+1I!(I)j=1I1(N+j2(I1))\displaystyle\frac{1}{I!}(N-2(I-1))\prod_{j=1}^{I-1}(N+j-2(I-1))+\frac{1}{I!}(I)\prod_{j=1}^{I-1}(N+j-2(I-1))
=\displaystyle= 1I!j=1I(N+j2(I1)),\displaystyle\frac{1}{I!}\prod_{j=1}^{I}(N+j-2(I-1)),

as desired. ∎

Remark 3.17.

We note from the proof of the theorem above a few facts for future reference. By our base cases, we have (R(0,t)^)=1+1=2\ell(\widehat{R(0,t)})=1+1=2 and (R(1,t)^)=1+2=3\ell(\widehat{R(1,t)})=1+2=3. Taking the Fibonacci sequence F(n)F(n) with F(0)=0F(0)=0 and F(1)=1F(1)=1, we have F(2)=1F(2)=1, F(3)=2F(3)=2, and F(4)=3F(4)=3, so that

(R(0,t)^)\displaystyle\ell(\widehat{R(0,t)}) =\displaystyle= F(3)\displaystyle F(3)
(R(1,t)^)\displaystyle\ell(\widehat{R(1,t)}) =\displaystyle= F(4).\displaystyle F(4).

These facts become useful in Proposition 3.19.

We see in the following corollary that the regularity of R(n,t)R(n,t) is n/2+1\left\lfloor n/2\right\rfloor+1. For an alternate proof of the regularity of R(n,t)R(n,t) which uses different machinery and more graph-theoretic properties, see the end of this section.

Corollary 3.18.

For GntG_{n}^{t}\in\mathcal{F},

regR(n,t)=n/2+1.\operatorname{reg}R(n,t)=\left\lfloor n/2\right\rfloor+1.
Proof.

We show that regR(n,t)\operatorname{reg}R(n,t) is equal to the top nonzero degree of R(n,t)^\widehat{R(n,t)} and that this value agrees with the above. Since R(n,t)^\widehat{R(n,t)} is Artinian by Remark 3.11, it is clear that regR(n,t)^\operatorname{reg}\widehat{R(n,t)} is the top nonzero degree of R(n,t)^\widehat{R(n,t)}. By Theorem 3.16, we know the top nonzero degree is NN for some Nn/2+1N\leq n/2+1, so that Nn/2+1N\leq\left\lfloor n/2\right\rfloor+1. In fact, the top nonzero degree is n/2+1\left\lfloor n/2\right\rfloor+1, provided dn,n/2+10d_{n,\left\lfloor n/2\right\rfloor+1}\neq 0. The only way we have a factor of zero in

dn,n/2+1=1(n/2+1)!j=1n/2+1(n+j2(n/2))d_{n,\lfloor n/2\rfloor+1}={\displaystyle\frac{1}{(\lfloor n/2\rfloor+1)!}\prod_{j=1}^{\lfloor n/2\rfloor+1}(n+j-2(\lfloor n/2\rfloor))}

is if n+j2(n/2)=0n+j-2(\lfloor n/2\rfloor)=0, which means j/2=n/2n/2j/2=\lfloor n/2\rfloor-n/2. This can only happen when j<1j<1, but j1j\geq 1, so we conclude that dn,n/2+10d_{n,\left\lfloor n/2\right\rfloor+1}\neq 0. Since R(n,t)^\widehat{R(n,t)} is an Artinian quotient of R(n,t)R(n,t) by a linear regular sequence, we conclude that

regR(n,t)=regR(n,t)^=n/2+1.\operatorname{reg}R(n,t)=\operatorname{reg}\widehat{R(n,t)}=\left\lfloor n/2\right\rfloor+1.

In the following, we first compute the lengths of the dimension zero rings R(n,t)^\widehat{R(n,t)}, and then show a closed form for the multiplicity of our original rings R(n,t)R(n,t) by using a Fibonacci relationship between the lengths of the rings R(n,t)^\widehat{R(n,t)} and applying Binet’s formula for F(n)F(n), the nnth number in the Fibonacci sequence:

F(n)=(1+5)n(15)n2n5.F(n)=\frac{(1+\sqrt{5})^{n}-(1-\sqrt{5})^{n}}{2^{n}\sqrt{5}}.

In the theorem and corollaries which follow, we suppress tt for convenience, since the statements are independent of tt.

Proposition 3.19.

The lengths of the rings R(n)^\widehat{R(n)} satisfy the recursive formula

(R(n)^)=(R(n1)^)+(R(n2)^)\ell(\widehat{R(n)})=\ell(\widehat{R(n-1)})+\ell(\widehat{R(n-2)})

for n2n\geq 2. Consequently, if F(n)F(n) is the Fibonacci sequence, with F(0)=0F(0)=0 and F(1)=1F(1)=1, then

(R(n)^)=F(n+3)=(1+5)n+3(15)n+32n+35.\ell(\widehat{R(n)})=F\left(n+3\right)=\frac{(1+\sqrt{5})^{n+3}-(1-\sqrt{5})^{n+3}}{2^{n+3}\sqrt{5}}.
Proof.

Again, we use Notation 3.14. By the recursive relationship from Lemma 3.15, since dn,0=1d_{n,0}=1 in general, and since dn,i=0d_{n,i}=0 in general for i>n/2+1i>n/2+1 by Theorem 3.16, we have for n2n\geq 2 that

(R(n)^)=i=0n/2+1dn,i\displaystyle\ell(\widehat{R(n)})=\sum_{i=0}^{\left\lfloor n/2\right\rfloor+1}d_{n,i} =\displaystyle= 1+i=1n/2+1(dn1,i+dn2,i1)\displaystyle 1+\sum_{i=1}^{\left\lfloor n/2\right\rfloor+1}\left(d_{n-1,i}+d_{n-2,i-1}\right)
=\displaystyle= i=0n/2+1dn1,i+i=0(n2)/2+1dn2,i\displaystyle\sum_{i=0}^{\left\lfloor n/2\right\rfloor+1}d_{n-1,i}+\sum_{i=0}^{\left\lfloor(n-2)/2\right\rfloor+1}d_{n-2,i}
=\displaystyle= (R(n1)^)+(R(n2)^).\displaystyle\ell(\widehat{R(n-1)})+\ell(\widehat{R(n-2)}).

Now we show the second statement. For our base cases, we see from Remark 3.17 that (R(0)^)=F(3)=F(0+3)\ell(\widehat{R(0)})=F(3)=F(0+3) and that (R(1)^)=F(4)=F(1+3)\ell(\widehat{R(1)})=F(4)=F(1+3).

Now suppose that (R(n1)^)=F(n+2){\displaystyle\ell(\widehat{R(n-1)})=F\left(n+2\right)} and (R(n2)^)=F(n+1){\displaystyle\ell(\widehat{R(n-2)})=F\left(n+1\right)}.
Then we have

(R(n)^)=(R(n1)^)+(R(n2)^)=F(n+3),\ell(\widehat{R(n)})=\ell(\widehat{R(n-1)})+\ell(\widehat{R(n-2)})\\ =F\left(n+3\right),

as desired. The closed form for (R(n)^)\ell(\widehat{R(n)}) follows directly from Binet’s formula for the Fibonacci sequence. ∎

Corollary 3.20.

For n2n\geq 2, there is an equality of multiplicities

e(R(n)^)=e(R(n1)^)+e(R(n2)^).e(\widehat{R(n)})=e(\widehat{R(n-1)})+e(\widehat{R(n-2)}).
Proof.

We have established the length of the Artinian rings R(n)^\widehat{R(n)}, and hence the multiplicity e(R(n)^)e(\widehat{R(n)}). ∎

Corollary 3.21.

For n2n\geq 2, there is an equality of multiplicities

e(R(n))=e(R(n1))+e(R(n2)).e(R(n))=e(R(n-1))+e(R(n-2)).

In particular,

e(R(n))=F(n+3)=(1+5)n+3(15)n+32n+35.e(R(n))=F\left(n+3\right)=\frac{(1+\sqrt{5})^{n+3}-(1-\sqrt{5})^{n+3}}{2^{n+3}\sqrt{5}}.
Proof.

To obtain the multiplicity of R(n)R(n), we look at R(n)^=R(n)/(Xn¯)\widehat{R(n)}=R(n)/(\overline{X_{n}}), which by Remark 3.11 and Corollary 3.12 is the Artinian quotient of R(n)R(n) by a linear regular sequence. By a standard result, we may calculate length along the obvious short exact sequences coming from multiplication by elements of our regular sequence to obtain the equality

HilbR(n)(t)(1t)d=HilbR(n)^(t),\text{Hilb}_{R(n)}(t)(1-t)^{d}=\text{Hilb}_{\widehat{R(n)}}(t),

where dd is the Krull dimension of R(n)R(n). Defining multiplicity as in and preceding [24, Thm 16.7], it follows immediately that

e(R(n))=HilbR(n)(t)(1t)d|t=1=HilbR(n)^(1)=(R(n)^).e(R(n))=\text{Hilb}_{R(n)}(t)(1-t)^{d}\big{\rvert}_{t=1}=\text{Hilb}_{\widehat{R(n)}}(1)=\ell(\widehat{R(n)}).

We are done by Proposition 3.19. ∎

We reintroduce tt and spend the remainder of this section providing an alternate
graph-theoretic proof for the regularity of R(n,t)R(n,t).

Alternate proof of Corollary 3.18.

We show that regR(n,t)=n/2+1\operatorname{reg}R(n,t)=\left\lfloor n/2\right\rfloor+1 by proving that regIGnt=n/2+2\operatorname{reg}I_{G_{n}^{t}}=\left\lfloor n/2\right\rfloor+2. We first show that

regIGntn/2+2.\operatorname{reg}I_{G_{n}^{t}}\leq\left\lfloor n/2\right\rfloor+2.

We recall by Proposition 2.13 that the graph GntG_{n}^{t} is chordal bipartite with vertex bipartition V1V2V_{1}\cup V_{2} of cardinalities

|V1|\displaystyle|V_{1}| =\displaystyle= n2+2\displaystyle\left\lfloor\frac{n}{2}\right\rfloor+2
|V2|\displaystyle|V_{2}| =\displaystyle= n2+2,\displaystyle\left\lceil\frac{n}{2}\right\rceil+2,

and that GntG_{n}^{t} does not have any vertices of degree one by Remark 2.6. Then by Theorem 4.9 of [3], we have

regIGntmin{n2+2,n2+2}=n2+2.\operatorname{reg}I_{G_{n}^{t}}\leq\min\left\{\left\lfloor\frac{n}{2}\right\rfloor+2,\left\lceil\frac{n}{2}\right\rceil+2\right\}=\left\lfloor\frac{n}{2}\right\rfloor+2.

We note that we may equivalently prove regR(n,t)n2+1\operatorname{reg}R(n,t)\leq\left\lfloor\frac{n}{2}\right\rfloor+1 by choosing the n2+2\left\lfloor\frac{n}{2}\right\rfloor+2 edges whose indices are equivalent to zero modulo 44, one from each row of LntL_{n}^{t}, to obtain an edge matching (different from an induced matching, below) and then applying [14, Th 1].

We now show that regIGntn/2+2\operatorname{reg}I_{G_{n}^{t}}\geq\left\lfloor n/2\right\rfloor+2. Since IGntI_{G_{n}^{t}} is homogeneous and in>IGnt\text{in}_{>}I_{G_{n}^{t}} consists of squarefree monomials by Corollary 3.3, we have by Corollary 2.7 of [4] that regin>IGnt=regIGnt\operatorname{reg}\text{in}_{>}I_{G_{n}^{t}}=\operatorname{reg}I_{G_{n}^{t}}, so it suffices to prove that regin>IGntn/2+2\operatorname{reg}\text{in}_{>}I_{G_{n}^{t}}\geq\left\lfloor n/2\right\rfloor+2. The ideal in>IGnt\text{in}_{>}I_{G_{n}^{t}} can be viewed as the edge ideal of a simple graph, a “comb” with n+1n+1 tines, with consecutive odd variables corresponding to vertices along the spine, as pictured below:

x2x_{2}x3x_{3}x5x_{5}x4x_{4}x7x_{7}x6x_{6}x9x_{9}x8x_{8}x11x_{11}x10x_{10}x2n+1x_{2n+1}x2nx_{2n}x2n+3x_{2n+3}x2n+2x_{2n+2}

We know from Theorem 6.5 of [13] that the regularity of an edge ideal is bounded below by the number of edges in any induced matching plus one, so we choose n/2+1\left\lfloor{n/2}\right\rfloor+1 edges (tines) corresponding to certain odd variables that create an induced matching. By beginning with the x3x_{3}-tine and choosing every other tine corresponding to the variables

x3,x3+4(1),,x3+4(n/2),x_{3},x_{3+4(1)},\ldots,x_{3+4(\left\lfloor{n/2}\right\rfloor)},

we obtain n/2+1\left\lfloor{n/2}\right\rfloor+1 edges that are an induced matching, so we have

regin>IGntn/2+2,\operatorname{reg}\text{in}_{>}I_{G_{n}^{t}}\geq\left\lfloor{n/2}\right\rfloor+2,

as desired.

We conclude that regIGnt=n/2+2\operatorname{reg}I_{G_{n}^{t}}=\left\lfloor n/2\right\rfloor+2, and hence that regR(n,t)=n/2+1.\operatorname{reg}R(n,t)=\left\lfloor n/2\right\rfloor+1.

References

  • [1] L. Ballard. Properties of the Toric Rings of a Chordal Bipartite Family of Graphs. PhD thesis, Syracuse University, 2020.
  • [2] S. K. Beyarslan, H. T. Hà, and A. O’Keefe. Algebraic properties of toric rings of graphs. Communications in Algebra, 47(1):1–16, 2019.
  • [3] J. Biermann, A. O’Keefe, and A. Van Tuyl. Bounds on the regularity of toric ideals of graphs. Advances in Applied Mathematics, 85:84–102, 2017.
  • [4] A. Conca and M. Varbaro. Square-free Gröbner degenerations. Inventiones Mathematicae, 221(3):713–730, 2020.
  • [5] A. Corso and U. Nagel. Monomial and toric ideals associated to Ferrers graphs. Transactions of the American Mathematical Society, 361(3):1371–1395, 2009.
  • [6] D. Cox, J. Little, and D. O’Shea. Ideals, varieties, and algorithms. Undergraduate Texts in Mathematics. Springer, New York, third edition, 2007. An introduction to computational algebraic geometry and commutative algebra.
  • [7] A. D’Alì. Toric ideals associated with gap-free graphs. Journal of Pure and Applied Algebra, 219(9):3862–3872, 2015.
  • [8] G. Favacchio, G. Keiper, and A. Van Tuyl. Regularity and h-polynomials of toric ideals of graphs. Proceedings of the American Mathematical Society, 148:4665–4677, 2020.
  • [9] F. Galetto, J. Hofscheier, G. Keiper, C. Kohne, M. E. Uribe Paczka, and A. Van Tuyl. Betti numbers of toric ideals of graphs: a case study. Journal of Algebra and its Applications, 18(12):1950226, 2019.
  • [10] I. Gitler and C. E. Valencia. Multiplicities of edge subrings. Discrete Mathematics, 302(1):107–123, 2005.
  • [11] D. R. Grayson and M. E. Stillman. Macaulay2, a software system for research in algebraic geometry. Available at http://www.math.uiuc.edu/Macaulay2/.
  • [12] Z. Greif and J. McCullough. Green-Lazarsfeld condition for toric edge ideals of bipartite graphs. Journal of Algebra, 562:1–27, 2020.
  • [13] H. T. Hà and A. Van Tuyl. Monomial ideals, edge ideals of hypergraphs, and their graded Betti numbers. Journal of Algebraic Combinatorics, 27(2):215–245, 2008.
  • [14] J. Herzog and T. Hibi. The regularity of edge rings and matching numbers. Mathematics, 8(1):39, 2020.
  • [15] J. Herzog, T. Hibi, and H. Ohsugi. Binomial ideals, volume 279 of Graduate Texts in Mathematics. Springer, Cham, 2018.
  • [16] T. Hibi, A. Higashitani, K. Kimura, and A. O’Keefe. Depth of edge rings arising from finite graphs. Proceedings of the American Mathematical Society, 139(11):3807–3813, 2011.
  • [17] T. Hibi and L. Katthän. Edge rings satisfying Serre’s condition (R1{R}_{1}). Proceedings of the American Mathematical Society, 142(7):2537–2541, 2014.
  • [18] T. Hibi, K. Matsuda, and H. Ohsugi. Strongly Koszul edge rings. Acta Mathematica Vietnamica, 41(1):69–76, 2016.
  • [19] T. Hibi, K. Matsuda, and A. Tsuchiya. Edge rings with 3-linear resolutions. Proceedings of the American Mathematical Society, 147(8):3225–3232, 2019.
  • [20] T. Hibi and H. Ohsugi. Toric ideals generated by quadratic binomials. Journal of Algebra, 218(2):509–527, 1999.
  • [21] M. Kreuzer and L. Robbiano. Computational Commutative Algebra 1. Springer, Berlin, Heidelberg, 2000.
  • [22] K. Mori, H. Ohsugi, and A. Tsuchiya. Edge rings with qq-linear resolutions. Preprint, arxiv.org/abs/2010.02854.
  • [23] R. Nandi and R. Nanduri. On Betti numbers of toric algebras of certain bipartite graphs. Journal of Algebra and Its Applications, 18(12):1950231, 2019.
  • [24] I. Peeva. Graded Syzygies. Springer London, London, 2011.
  • [25] A. Simis, W. V. Vasconcelos, and R. H. Villarreal. On the ideal theory of graphs. Journal of Algebra, 167(2):389–416, 1994.
  • [26] C. Tatakis and A. Thoma. On the universal Gröbner bases of toric ideals of graphs. Journal of Combinatorial Theory, Series A, 118(5):1540–1548, 2011.
  • [27] R. H. Villarreal. Rees algebras of edge ideals. Communications in Algebra, 23(9):3513–3524, 1995.