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

Generalized Weights of Codes over Rings
and Invariants of Monomial Ideals

Elisa Gorla and Alberto Ravagnani
Abstract

We develop an algebraic theory of supports for RR-linear codes of fixed length, where RR is a finite commutative unitary ring. A support naturally induces a notion of generalized weights and allows one to associate a monomial ideal to a code. Our main result states that, under suitable assumptions, the generalized weights of a code can be obtained from the graded Betti numbers of its associated monomial ideal. In the case of 𝔽q{\mathbb{F}}_{q}-linear codes endowed with the Hamming metric, the ideal coincides with the Stanley-Reisner ideal of the matroid associated to the code via its parity-check matrix. In this special setting, we recover the known result that the generalized weights of an 𝔽q{\mathbb{F}}_{q}-linear code can be obtained from the graded Betti numbers of the ideal of the matroid associated to the code. We also study subcodes and codewords of minimal support in a code, proving that a large class of RR-linear codes is generated by its codewords of minimal support.


Introduction

In the past seventy years, much effort has been devoted to the study of algebraic and combinatorial objects associated to linear error-correcting codes. Of particular interest is the matroid associated to a linear code via its parity-check matrix, whose circuits are the minimal Hamming supports of the codewords. Many central results in classical coding theory, including the celebrated MacWilliams identities, their generalizations, and the duality between puncturing and shortening can be elegantly obtained via this correspondence; see e.g. [2, 3, 4, 11] and the references therein.

The matroid associated to a linear code via its parity check matrix retains a wealth of information about the structure of the code, including its length, dimension, minimum distance, weight distribution, and generalized weights. Moreover, the weight enumerator is determined by the Tutte polynomial of the matroid, see [8]. In addition, in [10] it is shown that the code’s generalized weights are determined by the graded Betti numbers of the Stanley-Reisner ideal of the matroid. The approach of [10] heavily relies on matroid theory and on the properties of the Hamming support.

In this paper, we depart from the classical theory of linear codes over a finite field and consider instead RR-linear codes CRnC\subseteq R^{n}, where RR is a finite commutative unitary ring. We start by proposing a general definition of support as a function σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} that enjoys a few natural properties. This naturally extends the notion of Hamming support traditionally studied in coding theory [12, page 177]. We give several examples of supports and operations to construct new supports from old. We define the support of a code CRnC\subseteq R^{n} as the join of the supports of its elements.

We then define the generalized weights of a code via the supports of its subcodes. Moreover, we identify a class of supports under which the algebra of the module RnR^{n} is compatible with the combinatorics of the poset u\mathbb{N}^{u} with the product order. We call these supports modular and establish some of their structural properties. As one might expect, the Hamming support is an example of a modular support.

Most of the paper is devoted the study of codes CRnC\subseteq R^{n} endowed with modular supports. We characterize their minimal codewords and prove that, if RR is a principal ideal ring, then their generalized weights are attained by subcodes that are minimally generated by codewords of minimal support in CC. We also provide evidence in various examples that our results do not extend to support functions that are not modular.

The centerpiece of this paper is a result connecting the theory of modular supports with invariants of monomial ideals. We associate a monomial ideal to a code CRnC\subseteq R^{n} via the supports of its codewords. Under this correspondence, inclusion of supports translates into divisibility among monomials. In Theorem 4.4 we prove that, under suitable assumptions, the generalized weights of an RR-linear code endowed with a modular support are determined by the graded Betti numbers of the associated monomial ideal. This generalizes a result of [10], with a stand-alone proof that does not rely on matroid theory.

We conclude the paper with a series of observations on the Hamming support. We review known results in the light of our contribution, such as the fact that every 𝔽q\mathbb{F}_{q}-linear code is generated by its codewords of minimal support. We also show that the same is true for those of maximal support, provided that qq is sufficiently large (and false in general for binary codes).

Outline.

In Section 1 we define RR-linear codes and review some algebra results about finite commutative rings and finite local rings. In Section 2 we define (modular) supports and generalized weights, establishing their main properties. Codewords and subcodes of minimal supports are studied in Section 3. In Section 4 we associate a monomial ideal to a code. Moreover, we prove that the generalized weights of the code are determined by the graded Betti numbers of the corresponding ideal. We study the Hamming support in Section 5.

Acknowledgements.

We are grateful to Maria Evelina Rossi for suggesting Example 1.12.

1 Codes and minimal systems of generators

In this section we introduce codes over finite commutative rings and describe some properties of their systems of generators.

Notation 1.1.

Throughout the paper nn and mm denote positive integers and (R,+,)(R,+,\cdot) is a finite commutative ring. All rings in this paper are unitary with 101\neq 0. We denote by ={0,1,2,}{\mathbb{N}}=\{0,1,2,\ldots\} the set of natural numbers. For aa\in\mathbb{N} we let [a]:={1,,a}[a]:=\{1,\ldots,a\}.

A classical theorem in commutative algebra states that every finite commutative ring RR is isomorphic to a finite product of finite local rings. We forget the isomorphism and write

R=R1××R,R=R_{1}\times\ldots\times R_{\ell}, (1.1)

where R1,,RR_{1},\ldots,R_{\ell} are finite local rings. For i[]i\in[\ell], let 𝔐i\mathfrak{M}_{i} be the maximal ideal of RiR_{i} and let J=J(R)J=J(R) be the Jacobson radical of RR. Recall that the Jacobson radical of a commutative ring RR is the intersection of all maximal ideals of RR, equivalently

J={rR: 1+rs is invertible for every sR}.J=\{r\in R\,:\,1+rs\mbox{ is invertible for every }s\in R\}.

It is easy to check that in our situation

J=𝔐1××𝔐R1××R=R.J=\mathfrak{M}_{1}\times\ldots\times\mathfrak{M}_{\ell}\subseteq R_{1}\times\ldots\times R_{\ell}=R.

If RR is a finite principal ideal ring, then each RiR_{i} is a finite chain ring. For i[]i\in[\ell], let 𝔐i=(αi)\mathfrak{M}_{i}=(\alpha_{i}). Then J=(α)J=(\alpha), where α=(α1,,α)\alpha=(\alpha_{1},\ldots,\alpha_{\ell}). In particular, if RR is a product of finite fields, then 𝔐i=0\mathfrak{M}_{i}=0 for i[]i\in[\ell] and J=0J=0. If R=𝔽qR=\mathbb{F}_{q}, then =1\ell=1 and J=0J=0 is the only maximal ideal of RR.

We denote by (R,𝔐)(R,\mathfrak{M}) a local ring RR with maximal ideal 𝔐\mathfrak{M}. If (R,𝔐)(R,\mathfrak{M}) is a finite local ring, then R/𝔐R/\mathfrak{M} is a finite field.


In this paper we consider codes of fixed length over the alphabet RR. All of our codes are assumed to be linear over RR.

Definition 1.2.

An RR-linear code, or simply a code, is an RR-submodule CRnC\subseteq R^{n}. The elements of CC are called codewords. A subcode of CC is an RR-submodule DCD\subseteq C.

Denote by eie_{i} the element of RR which corresponds to (0,,0,1,0,,0)R1××R(0,\ldots,0,1,0,\ldots,0)\in R_{1}\times\ldots\times R_{\ell}, where the one appears in the iith component. From the decomposition in (1.1) one has

Rn=R1n××Rn.R^{n}=R_{1}^{n}\times\ldots\times R_{\ell}^{n}.

In the sequel, for i{1,,}i\in\{1,\ldots,\ell\} we denote by πi:RnRin\pi_{i}:R^{n}\rightarrow R_{i}^{n} the standard projection on the iith coordinate.

Let CRnC\subseteq R^{n} be a code. For any v=(v1,,v)Cv=(v_{1},\ldots,v_{\ell})\in C, with vi=πi(v)Rinv_{i}=\pi_{i}(v)\in R_{i}^{n}, one has

(0,,0,vi,0,,0)=eivC.(0,\ldots,0,v_{i},0,\ldots,0)=e_{i}v\in C.

Hence, up to isomorphism, CC can be uniquely written as

C=C1××CRn,C=C_{1}\times\ldots\times C_{\ell}\subseteq R^{n}, (1.2)

where Ci=πi(C)RinC_{i}=\pi_{i}(C)\subseteq R_{i}^{n} for all i[]i\in[\ell]. We often consider codes C0:RnJC\subseteq 0:_{R^{n}}J. Recall that

0:RnJ={vRn:rv=0 for all rJ}.0:_{R^{n}}J=\{v\in R^{n}\,:\,rv=0\mbox{ for all }r\in J\}.

Then

0:RnJ=(0:R1n𝔐1)××(0:Rn𝔐).0:_{R^{n}}J=\left(0:_{R_{1}^{n}}\mathfrak{M}_{1}\right)\times\ldots\times\left(0:_{R_{\ell}^{n}}\mathfrak{M}_{\ell}\right).

Since 0:RnJ0:_{R^{n}}J is an RR-module annihilated by JJ, it is an R/JR/J-module. Hence, if (R,𝔐)(R,\mathfrak{M}) is a local ring, then 0:Rn𝔐0:_{R^{n}}\mathfrak{M} is a vector space over R/𝔐R/\mathfrak{M}.

For CRnC\subseteq R^{n} a code, we also consider the socle

0:CJ={vC:rv=0 for all rJ}=C(0:RnJ).0:_{C}J=\{v\in C\,:\,rv=0\mbox{ for all }r\in J\}=C\cap(0:_{R^{n}}J).

The socle of CC is a the largest subcode of CC which is annihilated by JJ. In particular, if (R,𝔐)(R,\mathfrak{M}) is a local ring, then 0:C𝔐0:_{C}\mathfrak{M} is an R/𝔐R/\mathfrak{M}-vector space.

A minimal system of generators of a code CRnC\subseteq R^{n} is a subset of CC whose elements generate CC and which is minimal with respect to inclusion. Notice that any system of generators of a code CC contains a minimal system of generators of CC.

Definition 1.3.

We denote by μ(C)\mu(C) the least cardinality of a system of generators of a code CC, with μ(0)=0\mu(0)=0 by convention. For a code C=C1××CRnC=C_{1}\times\ldots\times C_{\ell}\subseteq R^{n} as in (1.2), let

M(C):=μ(C1)++μ(C).M(C):=\mu(C_{1})+\ldots+\mu(C_{\ell}).
Example 1.4.

Let R=R1××RR=R_{1}\times\dots\times R_{\ell}, with RiR_{i} a finite local ring for i[]i\in[\ell]. Then

M(Rn)=μ(R1n)++μ(Rn)=n.M(R^{n})=\mu(R_{1}^{n})+\ldots+\mu(R_{\ell}^{n})=n\ell.

Clearly μ(C)M(C)\mu(C)\leq M(C) for every code CRnC\subseteq R^{n}. If RR is a finite local ring, all minimal systems of generators of a code CRnC\subseteq R^{n} have the same cardinality μ(C)=M(C)\mu(C)=M(C). This is a consequence of the next lemma, which summarizes some well-known properties of systems of generators of modules over local rings; see e.g. [13, Theorem 2.3].

Theorem 1.5.

Let (R,𝔐)(R,\mathfrak{M}) be a local ring and let C=v1,,vtC=\langle v_{1},\ldots,v_{t}\rangle be an RR-module. The elements v1,,vtv_{1},\ldots,v_{t} are a minimal system of generators of CC if and only if the equivalence classes v1¯,,vt¯\overline{v_{1}},\ldots,\overline{v_{t}} are an R/𝔐R/\mathfrak{M}-basis of the vector space C/𝔐CC/\mathfrak{M}C. In particular, every minimal system of generators of CC has cardinality μ(C)=dimR/𝔐(C/𝔐C)\mu(C)=\dim_{R/\mathfrak{M}}(C/\mathfrak{M}C).

Over an arbitrary RR, however, not all minimal system of generators of a code CRnC\subseteq R^{n} have the same cardinality.

Example 1.6.

Let R=6R=\mathbb{Z}_{6} and D=(2,3)C=62D=\langle(2,3)\rangle\subseteq C=\mathbb{Z}_{6}^{2}. Then μ(D)=1\mu(D)=1. Moreover, (2,0)=4(2,3)D(2,0)=4(2,3)\in D and (0,3)=3(2,3)D(0,3)=3(2,3)\in D. Hence D=(2,0),(0,3)D=\langle(2,0),(0,3)\rangle and {(2,0),(0,3)}\{(2,0),(0,3)\} is a minimal system of generators of DD of cardinality 2=M(D)2=M(D).

Notation 1.7.

Let CRnC\subseteq R^{n} be a code and let

𝒮j(C):={DC subcode:D has a minimal system of generators of cardinality j}.\mathcal{S}_{j}(C):=\{D\subseteq C\mbox{ subcode}\,:\,\mbox{$D$ has a minimal system of generators of cardinality $j$}\}.

In particular, 𝒮j(Rn)\mathcal{S}_{j}(R^{n}) is the set of codes CRnC\subseteq R^{n} which have a minimal system of generators of cardinality jj.

One can show that M(C)M(C) is the largest cardinality of a minimal system of generators of CRnC\subseteq R^{n}.

Theorem 1.8.

If C𝒮i(Rn)C\in\mathcal{S}_{i}(R^{n}), then there exist v1,,viv_{1},\ldots,v_{i} minimal generators of CC with the property that

V={ejvk:(j,k)[]×[i],ejvk0}V=\{e_{j}v_{k}\,:\,(j,k)\in[\ell]\times[i],\;e_{j}v_{k}\neq 0\}

is a minimal system of generators of CC with |V|i|V|\geq i. Moreover

M(C)=max{i0:C𝒮i(Rn)}M(C)=\max\{i\geq 0\,:\,C\in\mathcal{S}_{i}(R^{n})\}

and any minimal system of generators of CC of cardinality M(C)M(C) has the same form as VV.

Proof.

Let w1,,wiw_{1},\ldots,w_{i} be a minimal system of generators of C=C1××CC=C_{1}\times\ldots\times C_{\ell}. Let Cj=0××0×Cj×0××0CC_{j}^{\prime}=0\times\ldots\times 0\times C_{j}\times 0\times\ldots\times 0\subseteq C. Observe that ejwkCe_{j}w_{k}\in C for all kk and jj and wk=e1wk++ewkw_{k}=e_{1}w_{k}+\ldots+e_{\ell}w_{k} for all k[i]k\in[i]. This proves that the set {ejwk:(j,k)[]×[i],ejwk0}\{e_{j}w_{k}\,:\,(j,k)\in[\ell]\times[i],\,e_{j}w_{k}\neq 0\} generates CC. Moreover, the set {ejwk:k[i],ejwk0}\{e_{j}w_{k}\,:\,k\in[i],\,e_{j}w_{k}\neq 0\} generates CjC_{j}^{\prime} for any j[]j\in[\ell].

Fix j[]j\in[\ell]. If ejw1,,ejwie_{j}w_{1},\ldots,e_{j}w_{i} do not form a minimal system of generators of CjC_{j}^{\prime}, suppose up to reindexing that ejw1,,ejwke_{j}w_{1},\ldots,e_{j}w_{k} do, for some k<ik<i. For h[i][k]h\in[i]\setminus[k], write ejwh=rh,1ejw1++rh,kejwke_{j}w_{h}=r_{h,1}e_{j}w_{1}+\ldots+r_{h,k}e_{j}w_{k} for some rh,1,,rh,kRr_{h,1},\ldots,r_{h,k}\in R. Let vh=whv_{h}=w_{h} for h[k]h\in[k], vh=whrh,1ejw1rh,kejwkv_{h}=w_{h}-r_{h,1}e_{j}w_{1}-\ldots-r_{h,k}e_{j}w_{k} for h[i][k]h\in[i]\setminus[k]. Then v1,,viv_{1},\ldots,v_{i} are a minimal system of generators of CC with the property that ejv1,,ejvke_{j}v_{1},\ldots,e_{j}v_{k} are a minimal system of generators of CjC_{j}^{\prime} and ejvk+1==ejvi=0e_{j}v_{k+1}=\ldots=e_{j}v_{i}=0. Notice that only the jjth coordinate of v1,,viv_{1},\ldots,v_{i} was affected by this operation, hence ehwk=ehvke_{h}w_{k}=e_{h}v_{k} for all k[i]k\in[i] if hjh\neq j. Performing this operation for all j[]j\in[\ell] produces a minimal system of generators v1,,viv_{1},\ldots,v_{i} of CC with the property that the set V={ejvk:(j,k)[]×[i],ejvk0}V=\{e_{j}v_{k}\,:\,(j,k)\in[\ell]\times[i],\,e_{j}v_{k}\neq 0\} is a minimal system of generators of CC. Moreover, for j[]j\in[\ell], {ejvk:k[i],ejvk0}\{e_{j}v_{k}\,:\,k\in[i],\,e_{j}v_{k}\neq 0\} is a minimal system of generators of CjC_{j}^{\prime}. Since v1,,vi0v_{1},\ldots,v_{i}\neq 0, for each k[i]k\in[i] there must be at least a j[]j\in[\ell] such that ejvk0e_{j}v_{k}\neq 0. This proves that |V|i|V|\geq i.

To prove the last part of the statement, let M=max{i0:C𝒮i(Rn)}.M=\max\{i\geq 0\,:\,C\in\mathcal{S}_{i}(R^{n})\}. Since CC has a minimal system of generators v1,,vMv_{1},\ldots,v_{M}, by the first part of the statement V={ejvk:(j,k)[]×[M],ejvk0}V=\{e_{j}v_{k}\,:\,(j,k)\in[\ell]\times[M],\,e_{j}v_{k}\neq 0\} is a minimal system of generators of CC of |V|=M|V|=M. Therefore, for each k[M]k\in[M] there is exactly one j[]j\in[\ell] with ejvk0e_{j}v_{k}\neq 0. Moreover, {vk:k[M],ejvk0}\{v_{k}\,:\,k\in[M],\,e_{j}v_{k}\neq 0\} is a minimal system of generators of CjC_{j}^{\prime} for j[]j\in[\ell], hence it has cardinality μ(Cj)\mu(C_{j}) by Theorem 1.5. It follows that M=μ(C1)++μ(C)=M(C)M=\mu(C_{1})+\ldots+\mu(C_{\ell})=M(C). ∎

We conclude the section with a few elementary properties of M(C)M(C).

Proposition 1.9.

Let RR be a finite commutative ring and let DC0:RnJD\subseteq C\subseteq 0:_{R^{n}}J be codes. Then

M(D)M(C)M(D)\leq M(C)

and equality holds if and only if D=CD=C.

Proof.

Write C=C1××CC=C_{1}\times\ldots\times C_{\ell} and D=D1××DD=D_{1}\times\ldots\times D_{\ell}. Since DCD\subseteq C, we have DiCi0:Rin𝔐iD_{i}\subseteq C_{i}\subseteq 0:_{R_{i}^{n}}\mathfrak{M}_{i} for all i[]i\in[\ell]. So CiC_{i} and DiD_{i} are Ri/𝔐iR_{i}/\mathfrak{M}_{i}-vector spaces and μ(Di)μ(Ci)\mu(D_{i})\leq\mu(C_{i}) for all i[]i\in[\ell] by Theorem 1.5. It follows that

M(D)=μ(D1)++μ(D)μ(C1)++μ(C)=M(C).M(D)=\mu(D_{1})+\ldots+\mu(D_{\ell})\leq\mu(C_{1})+\ldots+\mu(C_{\ell})=M(C).

Moreover, M(D)=M(C)M(D)=M(C) if and only if μ(Di)=μ(Ci)\mu(D_{i})=\mu(C_{i}) for all i[]i\in[\ell]. In this case, CiC_{i} and DiD_{i} are Ri/𝔐iR_{i}/\mathfrak{M}_{i}-vector spaces of the same dimension by Theorem 1.5. Hence they are equal, therefore D=CD=C. ∎

Notice that one may have DCRnD\subsetneq C\subseteq R^{n} with M(D)=M(C)M(D)=M(C). Some examples of this arise for instance from the fact that, over a principal ideal ring (PIR in the sequel), the value of M(C)M(C) does not change when replacing CC with its socle. We will use this fact repeatedly throughout the paper.

Proposition 1.10.

Let RR be a finite PIR and let CRnC\subseteq R^{n} be a code. Then

M(C)=M(0:CJ).M(C)=M(0:_{C}J).
Proof.

We may assume without loss of generality that RR is a finite chain ring. Indeed, if the result is true for finite chain rings, then write R=R1××RR=R_{1}\times\ldots\times R_{\ell} as a product of finite chain rings and C=C1××CC=C_{1}\times\ldots\times C_{\ell}, where CiRinC_{i}\subseteq R_{i}^{n} is a code for i[]i\in[\ell]. We have

M(C)=μ(C1)++μ(C)=μ(0:C1𝔐1)++μ(0:C𝔐)=M(0:CJ),M(C)=\mu(C_{1})+\ldots+\mu(C_{\ell})=\mu(0:_{C_{1}}\mathfrak{M}_{1})+\ldots+\mu(0:_{C_{\ell}}\mathfrak{M}_{\ell})=M(0:_{C}J),

where the last equality follows from

0:CJ=(0:C1𝔐1)××(0:C𝔐).0:_{C}J=(0:_{C_{1}}\mathfrak{M}_{1})\times\ldots\times(0:_{C_{\ell}}\mathfrak{M}_{\ell}).

In order to prove that μ(C)=μ(0:CJ)\mu(C)=\mu(0:_{C}J) for CRnC\subseteq R^{n} and RR a finite chain ring, observe that J=(α)J=(\alpha) is principal and

μ(C)=dimR/(α)(C/αC)=dimR/(α)(0:Cα)=μ(0:Cα),\mu(C)=\dim_{R/(\alpha)}(C/\alpha C)=\dim_{R/(\alpha)}(0:_{C}\alpha)=\mu(0:_{C}\alpha), (1.3)

where the first and last equalities follow from Theorem 1.5. The short exact sequence

00:CαCαC00\to 0:_{C}\alpha\to C\to\alpha C\to 0

induces an isomorphism C/αC0:CαC/\alpha C\cong 0:_{C}\alpha, which proves the central equality in (1.3). ∎

The statement of Proposition 1.10 also holds when C=RnC=R^{n} and RR is a product of finite Gorenstein local rings.

Example 1.11.

Write R=R1××RR=R_{1}\times\ldots\times R_{\ell} and suppose that each RiR_{i} is a finite Gorenstein local ring. Suppose first that =1\ell=1, i.e. RR is a Gorenstein local ring with maximal ideal 𝔐\mathfrak{M}. We have the following isomorphisms of R/𝔐R/\mathfrak{M}-vector spaces

Rn/𝔐Rn(R/𝔐R)n(0:R𝔐)n=0:Rn𝔐,R^{n}/\mathfrak{M}R^{n}\cong(R/\mathfrak{M}R)^{n}\cong(0:_{R}\mathfrak{M})^{n}=0:_{R^{n}}\mathfrak{M},

where the central isomorphism follows from the definition of a Gorenstein local ring. Then μ(Rn)=μ(0:Rn𝔐)=n\mu(R^{n})=\mu(0:_{R^{n}}\mathfrak{M})=n by Theorem 1.5.

For general \ell, one has

M(Rn)=μ(R1n)++μ(Rn).M(R^{n})=\mu(R_{1}^{n})+\ldots+\mu(R_{\ell}^{n}).

Moreover, 0:RnJ=(0:R1n𝔐1)××(0:Rn𝔐)0:_{R^{n}}J=\left(0:_{R_{1}^{n}}\mathfrak{M}_{1}\right)\times\ldots\times\left(0:_{R_{\ell}^{n}}\mathfrak{M}_{\ell}\right), hence

M(0:RnJ)=μ(0:R1n𝔐1)++μ(0:Rn𝔐).M(0:_{R^{n}}J)=\mu\left(0:_{R_{1}^{n}}\mathfrak{M}_{1}\right)+\ldots+\mu\left(0:_{R_{\ell}^{n}}\mathfrak{M}_{\ell}\right).

It follows from the previous case (=1\ell=1) that

μ(Rin)=μ(0:Rin𝔐)=n\mu(R_{i}^{n})=\mu(0:_{R_{i}^{n}}\mathfrak{M})=n

for all i[].i\in[\ell].

The argument of Example 1.11 also shows that, if the equality in Proposition 1.10 holds for C=RnC=R^{n}, for RR a finite local ring, then RR must be Gorenstein. However, Proposition 1.10 is not true in general over finite Gorenstein local rings. The next example illustrating this was suggested to us by Maria Evelina Rossi.

Example 1.12.

Let R=𝔽2[x,y]/(x2,y2)R=\mathbb{F}_{2}[x,y]/(x^{2},y^{2}). Then RR is a finite local ring with maximal ideal 𝔐=(x,y)\mathfrak{M}=(x,y). Let C=𝔐C=\mathfrak{M}. Then μ(C)=2\mu(C)=2, but μ(0:C𝔐)=μ(xy)=1\mu(0:_{C}\mathfrak{M})=\mu(\langle xy\rangle)=1.

2 Supports and generalized weights

In this section we develop an algebraic theory of supports over a finite commutative ring RR. We propose a general definition of support as a map RnuR^{n}\to\mathbb{N}^{u}, which naturally induces a notion of generalized weights for codes CRnC\subseteq R^{n}. This extends the notion of generalized Hamming weights for codes that are linear over a finite field 𝔽q{\mathbb{F}}_{q}. We establish some properties of support functions and generalized weights. We also define a family of supports, the modular supports, whose associated generalized weights will be studied in the next sections.

Notation 2.1.

In the sequel, u1u\geq 1 is an integer. For s,tus,t\in{\mathbb{N}}^{u} write sts\leq t if sitis_{i}\leq t_{i} for all i[u]i\in[u]. Then (u,)({\mathbb{N}}^{u},\leq) is a (poset) lattice. The meet of s,tus,t\in{\mathbb{N}}^{u} is the element stus\wedge t\in{\mathbb{N}}^{u} given by (st)i=min{si,ti}(s\wedge t)_{i}=\min\{s_{i},t_{i}\} for all i[u]i\in[u]. The join of s,tus,t\in{\mathbb{N}}^{u}, denoted by sts\vee t, is (st)i=max{si,ti}(s\vee t)_{i}=\max\{s_{i},t_{i}\} for all i[u]i\in[u]. For sus\in\mathbb{N}^{u}, we let |s|:=s1++su|s|:=s_{1}+\cdots+s_{u}.

Definition 2.2.

A support on RnR^{n} is a function σ:Rnu\sigma:R^{n}\to{\mathbb{N}}^{u} with the following properties.

  1. (P1)

    σ(v)=0\sigma(v)=0 if and only if v=0v=0.

  2. (P2)

    σ(rv)σ(v)\sigma(rv)\leq\sigma(v) for all rRr\in R and vRnv\in R^{n}.

  3. (P3)

    σ(v+w)σ(v)σ(w)\sigma(v+w)\leq\sigma(v)\vee\sigma(w) for all v,wRnv,w\in R^{n}.

A support function σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} satisfies the following additional properties.

Lemma 2.3.

Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} be a support. The following hold.

  1. (P4)

    If vRnv\in R^{n} and rRr\in R is a unit, then σ(rv)=σ(v)\sigma(rv)=\sigma(v).

  2. (P5)

    If v,wRnv,w\in R^{n} and i[u]i\in[u] satisfy σ(v)i=0\sigma(v)_{i}=0 and σ(w)i0\sigma(w)_{i}\neq 0, then σ(w+v)i0\sigma(w+v)_{i}\neq 0.

  3. (P6)

    If v,wRnv,w\in R^{n} and i[u]i\in[u] satisfy σ(v)i=σ(w)i=0\sigma(v)_{i}=\sigma(w)_{i}=0, then σ(w+v)i=0\sigma(w+v)_{i}=0.

Proof.

The first claim easily follows from Property (P2). To see the second, suppose towards a contradiction that σ(w+v)i=0\sigma(w+v)_{i}=0. Since w=(w+v)+(v)w=(w+v)+(-v), by (P3) and the first claim we have σ(w)σ(w+v)σ(v)=σ(w+v)σ(v)\sigma(w)\leq\sigma(w+v)\vee\sigma(-v)=\sigma(w+v)\vee\sigma(v), from which σ(w)i=0\sigma(w)_{i}=0, a contradiction. Finally, the third claim follows from Property (P3). ∎

A support σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} naturally induces a weight function wt:Rn\textnormal{wt}:R^{n}\to\mathbb{N}.

Definition 2.4.

The weight of vRnv\in R^{n} is wt(v):=|σ(v)|\textnormal{wt}(v):=|\sigma(v)|. The minimum weight and maximum weight of a code 0CRn0\neq C\subseteq R^{n} are, respectively,

minwt(C):=min{wt(v):vC{0}}andmaxwt(C):=max{wt(v):vC}.\min\textnormal{wt}(C):=\min\{\textnormal{wt}(v)\,:\,v\in C\setminus\{0\}\}\qquad\mbox{and}\qquad\max\textnormal{wt}(C):=\max\{\textnormal{wt}(v)\,:\,v\in C\}.

Notice that the function wt:Rn\textnormal{wt}:R^{n}\to{\mathbb{N}} has indeed the properties of a weight, since wt(v)0\textnormal{wt}(v)\geq 0 for all vv, wt(v)=0\textnormal{wt}(v)=0 if and only if v=0v=0, and wt(u+v)wt(u)+wt(v)\textnormal{wt}(u+v)\leq\textnormal{wt}(u)+\textnormal{wt}(v) for all u,vRnu,v\in R^{n}. In addition, the weight satisfies wt(rv)wt(v)\textnormal{wt}(rv)\leq\textnormal{wt}(v) for all rRr\in R and vRnv\in R^{n} and wt(rv)=wt(v)\textnormal{wt}(rv)=\textnormal{wt}(v) for rRr\in R invertible and vRnv\in R^{n}.

We give some examples of support functions. Many others can be obtained by applying Proposition 2.7 to these examples.

Example 2.5.
  1. (1)

    The function σ:𝔽22{0,1,2}3\sigma:\mathbb{F}_{2}^{2}\to\{0,1,2\}^{3} defined by σ(0,0)=(0,0,0)\sigma(0,0)=(0,0,0), σ(1,0)=(2,0,2)\sigma(1,0)=(2,0,2), σ(0,1)=(2,1,0)\sigma(0,1)=(2,1,0), and σ(1,1)=(0,1,2)\sigma(1,1)=(0,1,2) is a support.

  2. (2)

    Let RR a be finite ring and let 0=I0I1Iϵ1Iϵ=R0=I_{0}\subsetneq I_{1}\subsetneq\cdots\subsetneq I_{\epsilon-1}\subsetneq I_{\epsilon}=R be a chain of ideals of RR. For rRr\in R, let σ(r):=min{0iϵ:rIi}\sigma(r):=\min\{0\leq i\leq\epsilon\,:\,r\in I_{i}\}. Extend σ\sigma coordinatewise to σ:Rn{0,,ϵ}n\sigma:R^{n}\to\{0,\ldots,\epsilon\}^{n}. It can be checked that σ\sigma is a support, called the chain support on RR, see [15, Example 26].

  3. (3)

    Let RR be a finite chain ring. The chain support associated to the full chain of ideals of RR is the chain ring support on RR.

  4. (4)

    If R=𝔽qR=\mathbb{F}_{q} is a finite field, the chain ring support coincides with the Hamming support σH:𝔽qn{0,1}n\sigma^{\textnormal{H}}:\mathbb{F}_{q}^{n}\to\{0,1\}^{n}, given by σH(v)i=1\sigma^{\textnormal{H}}(v)_{i}=1 if vi0v_{i}\neq 0 and σH(v)i=0\sigma^{\textnormal{H}}(v)_{i}=0 if vi=0v_{i}=0. See [12] for a general reference on Hamming-metric codes.

  5. (5)

    In his master thesis [5], written under the direction of J. Rosenthal and V. Weger, N. Gassner introduces the pp-adic weight and distance on pen\mathbb{Z}_{p^{e}}^{n}, where pp is prime and e1e\geq 1. The pp-adic weight on pe\mathbb{Z}_{p^{e}} induces the same partition as the weight associated to the chain ring support of pe\mathbb{Z}_{p^{e}}.

Not all the supports studied in the coding theory literature are supports according to Definition 2.2.

Example 2.6.

The Lee weight wtL:4{0,1,2}\textnormal{wt}^{\textnormal{L}}:\mathbb{Z}_{4}\to\{0,1,2\} is defined by wtL(0)=0\textnormal{wt}^{\textnormal{L}}(0)=0, wtL(1)=wtL(3)=1\textnormal{wt}^{\textnormal{L}}(1)=\textnormal{wt}^{\textnormal{L}}(3)=1 and wtL(2)=2\textnormal{wt}^{\textnormal{L}}(2)=2. Its coordinatewise extension to 4n\mathbb{Z}_{4}^{n} is not a support in the sense of Definition 2.2. For instance, wtL(1+1)=2max{wtL(1),wtL(1)}=1\textnormal{wt}^{\textnormal{L}}(1+1)=2\not\leq\max\{\textnormal{wt}^{\textnormal{L}}(1),\textnormal{wt}^{\textnormal{L}}(1)\}=1, contradicting Property (P3).

In the next proposition we list some simple operations that allow one to construct new supports from known ones. The proof is straightforward and left to the reader.

Proposition 2.7.
  1. (1)

    Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} be a function and let s:uus:\mathbb{N}^{u}\to\mathbb{N}^{u} be a permutation of the coordinates. Then σ\sigma is a support if and only if sσs\circ\sigma is a support.

  2. (2)

    Let σi:Rniui\sigma_{i}:R^{n_{i}}\to\mathbb{N}^{u_{i}} be functions, i[]i\in[\ell]. Let n=n1++nn=n_{1}+\ldots+n_{\ell} and u=u1++uu=u_{1}+\ldots+u_{\ell}. Let

    σ=σ1××σ:Rnu,(v1,,v)(σ1(v1),,σ(v)).\sigma=\sigma_{1}\times\cdots\times\sigma_{\ell}:R^{n}\to\mathbb{N}^{u},\qquad(v_{1},\ldots,v_{\ell})\mapsto(\sigma_{1}(v_{1}),\ldots,\sigma_{\ell}(v_{\ell})).

    Then σ\sigma is a support if and only if σ1,,σ\sigma_{1},\ldots,\sigma_{\ell} are supports.

  3. (3)

    Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u}, σ(v)=(σ1(v),,σu(v))\sigma(v)=(\sigma_{1}(v),\ldots,\sigma_{u}(v)), be a function. Let i[u]i\in[u], a{0}a\in\mathbb{N}\setminus\{0\}, and define

    σi,a:Rnu,v(σ1(v),,σi1(v),aσi(v),σi+1(v),,σu(v)).\sigma_{i,a}:R^{n}\to\mathbb{N}^{u},\qquad v\mapsto(\sigma_{1}(v),\ldots,\sigma_{i-1}(v),a\sigma_{i}(v),\sigma_{i+1}(v),\ldots,\sigma_{u}(v)).

    Then σ\sigma is a support if and only if σi,a\sigma_{i,a} is.

  4. (4)

    Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u}, σ(v)=(σ1(v),,σu(v))\sigma(v)=(\sigma_{1}(v),\ldots,\sigma_{u}(v)). For i[u]i\in[u], let

    σ~i:Rnu+1,v(σ1(v),,σi1(v),σi(v),σi(v),σi+1(v),,σu(v)).\tilde{\sigma}_{i}:R^{n}\to\mathbb{N}^{u+1},\qquad v\mapsto(\sigma_{1}(v),\ldots,\sigma_{i-1}(v),\sigma_{i}(v),\sigma_{i}(v),\sigma_{i+1}(v),\ldots,\sigma_{u}(v)).

    Then σ\sigma is a support if and only if σ~i\tilde{\sigma}_{i} is.

  5. (5)

    Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u}, σ(v)=(σ1(v),,σu(v))\sigma(v)=(\sigma_{1}(v),\ldots,\sigma_{u}(v)), be a support and let i[u]i\in[u]. Assume that there are no a{0}a\in\mathbb{N}\setminus\{0\} and vRnv\in R^{n} such that σ(v)=(0,,0,a,0,,0)\sigma(v)=(0,\ldots,0,a,0,\ldots,0), where aa appears in the iith entry. Then

    σ^i:Rnu1,v(σ1(v),,σi1(v),σi+1(v),,σu(v))\hat{\sigma}_{i}:R^{n}\to\mathbb{N}^{u-1},\qquad v\mapsto(\sigma_{1}(v),\ldots,\sigma_{i-1}(v),\sigma_{i+1}(v),\ldots,\sigma_{u}(v))

    is a support.

  6. (6)

    Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} be a function, σ(v)=(σ1(v),,σu(v))\sigma(v)=(\sigma_{1}(v),\ldots,\sigma_{u}(v)). For i[u]i\in[u], let

    σˇi:Rnu+1,v(σ1(v),,σi1(v),σi(v),0,σi+1(v),,σu(v)).\check{\sigma}_{i}:R^{n}\to\mathbb{N}^{u+1},\qquad v\mapsto(\sigma_{1}(v),\ldots,\sigma_{i-1}(v),\sigma_{i}(v),0,\sigma_{i+1}(v),\ldots,\sigma_{u}(v)).

    Then σ\sigma is a support if and only if σˇi\check{\sigma}_{i} is.

  7. (7)

    Let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} be a support and let f:RkRnf:R^{k}\to R^{n} be an injective RR-linear map. Then σf:Rku\sigma\circ f:R^{k}\to\mathbb{N}^{u} is a support.

  8. (8)

    Let σi:Rnui\sigma_{i}:R^{n}\to\mathbb{N}^{u_{i}} be supports, i[k]i\in[k]. Then σ=(σ1,,σk):Rnu1++uk\sigma=(\sigma_{1},\ldots,\sigma_{k}):R^{n}\to\mathbb{N}^{u_{1}+\ldots+u_{k}} is a support.

Similarly to the situation of linear codes endowed with the Hamming support, a support function over a finite commutative ring RR induces a notion of support of a code. In turn, this allows us to define generalized weights for RR-linear codes.

Definition 2.8.

The support of a code CRnC\subseteq R^{n} is

σ(C):=vCσ(v)u.\sigma(C):=\bigvee_{v\in C}\sigma(v)\in{\mathbb{N}}^{u}.

Notice that the support of a code is determined by the supports of the vectors in any system of generators. More precisely, let C=v1,,vkRnC=\langle v_{1},\ldots,v_{k}\rangle\subseteq R^{n}. Since, by Definition 2.2,

σ(r1v1++rkvk)σ(r1v1)σ(rkvk)σ(v1)σ(vk)\sigma(r_{1}v_{1}+\ldots+r_{k}v_{k})\leq\sigma(r_{1}v_{1})\vee\ldots\vee\sigma(r_{k}v_{k})\leq\sigma(v_{1})\vee\ldots\vee\sigma(v_{k})

for any r1,,rkRr_{1},\ldots,r_{k}\in R, we have

σ(C)=i=1kσ(vi).\sigma(C)=\bigvee_{i=1}^{k}\sigma(v_{i}). (2.1)

Moreover, if DCD\subseteq C, then by definition σ(D)σ(C)\sigma(D)\leq\sigma(C).

Example 2.9.

Equation (2.1) does not hold for the Lee weight wtL:4{0,1,2}\textnormal{wt}^{\textnormal{L}}:\mathbb{Z}_{4}\to\{0,1,2\}, nor for its coordinatewise extension to 4n\mathbb{Z}_{4}^{n}. For example, wtL(1)=2>1=wtL(1)\textnormal{wt}^{\textnormal{L}}(\langle 1\rangle)=2>1=\textnormal{wt}^{\textnormal{L}}(1), see also Example 2.6.

Definition 2.10.

For r[M(C)]r\in[M(C)], the rr-th generalized weight of CC is the integer

dr(C):=min{|σ(D)|:D𝒮j(C) for some jr}.d_{r}(C):=\min\{|\sigma(D)|\,:\,D\in\mathcal{S}_{j}(C)\mbox{ for some }j\geq r\}.

We also set

d0(C):=0.d_{0}(C):=0.

It follows from Theorem 1.8 that for r[M(C)]r\in[M(C)] we have 𝒮r(C)\mathcal{S}_{r}(C)\neq\emptyset. Hence dr(C)d_{r}(C) is well-defined.

Remark 2.11.

For r[M(C)]r\in[M(C)] one has

dr(C)=min{|σ(D)|:D𝒮r(C)}.d_{r}(C)=\min\{|\sigma(D)|\,:\,D\in\mathcal{S}_{r}(C)\}.

Indeed, let jrj\geq r and let D𝒮j(C)D\in\mathcal{S}_{j}(C). Then there exists a DDD^{\prime}\subseteq D such that D𝒮r(C)D^{\prime}\in\mathcal{S}_{r}(C) and |σ(D)||σ(D)||\sigma(D^{\prime})|\leq|\sigma(D)|.

In the next lemma we collect a few easy consequences of Definition 2.10.

Lemma 2.12.

Let DCRnD\subseteq C\subseteq R^{n} be codes. The following hold.

  1. (1)

    d1(C)=minwt(C)d_{1}(C)=\min\textnormal{wt}(C).

  2. (2)

    dr(D)dr(C)d_{r}(D)\geq d_{r}(C) for r[min{M(D),M(C)}]r\in[\min\{M(D),M(C)\}].

  3. (3)

    dr+1(C)dr(C)d_{r+1}(C)\geq d_{r}(C) for r[M(C)1]r\in[M(C)-1].

  4. (4)

    dr(C)=min{|σ(D)|:DC,M(D)r}d_{r}(C)=\min\{|\sigma(D)|\,:\,D\subseteq C,\,M(D)\geq r\} for r[M(C)]r\in[M(C)].

Proof.

By Property (P2) one has σ(v)=σ(v)\sigma(\langle v\rangle)=\sigma(v) for any vCv\in C. Hence (1) follows, thanks to Remark 2.11. Part (2) holds since every subcode of DD is also a subcode of CC. Part (3) follows from observing that drd_{r} is the minimum of the function |σ(D)||\sigma(D)| for DD ranging over the set 𝒮r(C)𝒮M(C)(C)\mathcal{S}_{r}(C)\cup\ldots\cup\mathcal{S}_{M(C)}(C) and passing from rr to r+1r+1 we minimize over a subset. In order to prove (4), let iri\geq r and D𝒮i(C)D\in\mathcal{S}_{i}(C). Then M(D)iM(D)\geq i by Theorem 1.8. Therefore, if D𝒮i(C)D\in\mathcal{S}_{i}(C) for some iri\geq r, then D𝒮M(D)(C)D\in\mathcal{S}_{M(D)}(C) and M(D)rM(D)\geq r. Since every DCD\subseteq C belongs to 𝒮M(D)(C)\mathcal{S}_{M(D)}(C), then {D𝒮i(C) for some ir}={DC,M(D)r}\{D\in\mathcal{S}_{i}(C)\mbox{ for some }i\geq r\}=\{D\subseteq C,\,M(D)\geq r\}. Therefore

dr(C)=min{|σ(D)|:D𝒮i(C) for some ir}=min{|σ(D)|:DC,M(D)r}.d_{r}(C)=\min\{|\sigma(D)|\,:\,D\in\mathcal{S}_{i}(C)\mbox{ for some }i\geq r\}=\min\{|\sigma(D)|\,:\,D\subseteq C,\,M(D)\geq r\}.\qed

We now show how the structure of supports relate to the decomposition of RR in (1.1).

Proposition 2.13.

Let σ:Rnu\sigma:R^{n}\rightarrow\mathbb{N}^{u} be a support. Then for any v=(v1,,v)Rn=R1n××Rnv=(v_{1},\ldots,v_{\ell})\in R^{n}=R_{1}^{n}\times\ldots\times R_{\ell}^{n} we have

σ(v)=σ1(v1)σ(v),\sigma(v)=\sigma_{1}(v_{1})\vee\ldots\vee\sigma_{\ell}(v_{\ell}),

where σi:Rinu\sigma_{i}:R_{i}^{n}\rightarrow\mathbb{N}^{u} is as support defined via σi(vi):=σ(eiv)\sigma_{i}(v_{i}):=\sigma(e_{i}v) for all i[]i\in[\ell].

Proof.

It is easy to check that σi\sigma_{i} is well-defined and is a support for all i[]i\in[\ell]. One has σi(vi)=σ(eiv)σ(v)\sigma_{i}(v_{i})=\sigma(e_{i}v)\leq\sigma(v), hence σ1(v1)σ(v)σ(v)\sigma_{1}(v_{1})\vee\ldots\vee\sigma_{\ell}(v_{\ell})\leq\sigma(v). Furthermore,

σ(v)=σ(i=1eiv)i=1σ(eiv)=i=1σi(vi).\sigma(v)=\sigma\left(\sum_{i=1}^{\ell}e_{i}v\right)\leq\bigvee_{i=1}^{\ell}\sigma(e_{i}v)=\bigvee_{i=1}^{\ell}\sigma_{i}(v_{i}).

It follows that σ(v)=σ1(v1)σ(v)\sigma(v)=\sigma_{1}(v_{1})\vee\ldots\vee\sigma_{\ell}(v_{\ell}), as desired. ∎

Remark 2.14.

When RR is a finite chain ring, support functions on RR have a simple description. To see this, let α\alpha be a generator of the maximal ideal of RR and let ϵ=min{i>0:αi=0}\epsilon=\min\{i>0\,:\,\alpha^{i}=0\}. Let σ,τ:Ru\sigma,\tau:R\rightarrow\mathbb{N}^{u} be supports. Then σ=τ\sigma=\tau if and only if σ(αi)=τ(αi)\sigma(\alpha^{i})=\tau(\alpha^{i}) for 0iϵ10\leq i\leq\epsilon-1. Indeed, every element of R{0}R\setminus\{0\} is of the form rαir\alpha^{i} where rr is a unit and 0iϵ10\leq i\leq\epsilon-1, and σ(rαi)=σ(αi)\sigma(r\alpha^{i})=\sigma(\alpha^{i}). Therefore, a support σ:Ru\sigma:R\rightarrow\mathbb{N}^{u} corresponds to a set of vectors a(0),a(1),,a(ϵ1)ua^{(0)},a^{(1)},\ldots,a^{(\epsilon-1)}\in\mathbb{N}^{u} with the property that a(0)a(1)a(ϵ1)a^{(0)}\geq a^{(1)}\geq\ldots\geq a^{(\epsilon-1)}. The correspondence is determined by setting σ(αi)=a(i)\sigma(\alpha^{i})=a^{(i)} for all i{0,,ϵ1}i\in\{0,\ldots,\epsilon-1\}. In particular, any support on RR induces the same partition as a chain support.

2.1 Modular Supports

In this subsection we define and study a class of supports whose structure is closely related to the RR-module structure of RnR^{n}, and that we therefore call modular. This paper is primarily devoted to the study of generalized weights associated to modular supports.

Definition 2.15.

A support σ\sigma is modular if it satisfies the following:

  1. (P\star)

    If v,wRnv,w\in R^{n} and i[u]i\in[u] satisfy 0σ(v)iσ(w)i0\neq\sigma(v)_{i}\leq\sigma(w)_{i}, then there exists rRr\in R such that σ(v+rw)i<σ(v)i\sigma(v+rw)_{i}<\sigma(v)_{i}.

Remark 2.16.

By repeatedly applying Property (P\star), one obtains the following equivalent property:  If v,wRnv,w\in R^{n} and i[u]i\in[u] satisfy 0σ(v)iσ(w)i0\neq\sigma(v)_{i}\leq\sigma(w)_{i}, then there exists rRr\in R such that σ(v+rw)i=0\sigma(v+rw)_{i}=0.

As for supports, one can easily produce new modular supports from known ones.

Proposition 2.17.

Let σ:Rnu\sigma:R^{n}\rightarrow{\mathbb{N}}^{u} be a support. Following the notation and numbering of Proposition 2.7, we have:

  1. (1)

    σ\sigma is modular if and only if sσs\circ\sigma is modular;

  2. (2)

    σ1,,σ\sigma_{1},\ldots,\sigma_{\ell} are modular if and only if σ\sigma is modular;

  3. (3)

    σ\sigma is modular if and only if σi,a\sigma_{i,a} is modular;

  4. (4)

    σ\sigma is modular if and only if σ~i\tilde{\sigma}_{i} is modular;

  5. (5)

    if σ\sigma is modular, then σ^i\hat{\sigma}_{i} is modular;

  6. (6)

    σ\sigma is modular if and only if σˇi\check{\sigma}_{i} is modular;

  7. (7)

    if σ\sigma is modular, then σf\sigma\circ f is modular;

  8. (8)

    if σ1,,σk\sigma_{1},\ldots,\sigma_{k} are modular, then σ=(σ1,,σk)\sigma=(\sigma_{1},\ldots,\sigma_{k}) is modular.

Several, but not all, of the supports that we have encountered so far are modular.

Example 2.18.

Support (1) of Example 2.5 is modular, while an arbitrary chain support is not. Over a finite chain ring, the only modular chain support is the chain ring support. For example, the chain support on 4\mathbb{Z}_{4} associated with the chain 040\subsetneq\mathbb{Z}_{4} is not modular. Indeed, σ(2)=σ(4)=1\sigma(2)=\sigma(4)=1, but there is no r4r\in\mathbb{Z}_{4} with 24r=02-4r=0.

The Hamming support is an example of modular support.

Example 2.19.

It is easy to check that the chain ring support of Example 2.5(3) is modular. Hence a product of chain ring supports is modular by Proposition 2.7 and Proposition 2.17(2). In particular, the Hamming support is modular.

Example 2.20.

The supports of Remark 2.14 are modular if and only if (a(j))i(a(k))i(a^{(j)})_{i}\neq(a^{(k)})_{i} for all j,k{0,,ϵ1}j,k\in\{0,\ldots,\epsilon-1\} distinct and i[u]i\in[u].

We now give more examples of supports which are not modular.

Example 2.21.

Let R=𝔽2R=\mathbb{F}_{2} and let σ:𝔽22{0,1}2\sigma:\mathbb{F}_{2}^{2}\to\{0,1\}^{2} be defined by

σ(0,0)=(0,0),σ(1,0)=(1,1),σ(0,1)=(0,1),σ(1,1)=(1,1).\sigma(0,0)=(0,0),\quad\sigma(1,0)=(1,1),\quad\sigma(0,1)=(0,1),\quad\sigma(1,1)=(1,1).

Then σ\sigma is a support which is not modular.

Example 2.22.

The chain support on 6\mathbb{Z}_{6} associated with the chain (0)(2)6(0)\subsetneq(2)\subsetneq\mathbb{Z}_{6} is not modular. Indeed, σ(2)=1\sigma(2)=1 and σ(3)=2\sigma(3)=2, but there is no r6r\in\mathbb{Z}_{6} with 23r=02-3r=0.

The next result shows that every modular support over a finite commutative ring decomposes as a product of modular supports over finite local rings.

Theorem 2.23.

Let RR be a finite commutative ring and let σ:Rnu\sigma:R^{n}\rightarrow\mathbb{N}^{u} be a modular support. Up to a permutation of the coordinates of u\mathbb{N}^{u} we have σ=σ1××σ\sigma=\sigma_{1}\times\ldots\times\sigma_{\ell} where σi:Rinui\sigma_{i}:R_{i}^{n}\rightarrow\mathbb{N}^{u_{i}} for i[]i\in[\ell] and u1,,uu_{1},...,u_{\ell} are integers with u1++u=uu_{1}+\ldots+u_{\ell}=u. Moreover, σi\sigma_{i} is a modular support for all i[]i\in[\ell].

Proof.

For vRnv\in R^{n}, write v=(v1,,v)v=(v_{1},\ldots,v_{\ell}) with viRinv_{i}\in R_{i}^{n}. By Proposition 2.13 we have

σ(v1,,v)=σ1(v1)σ(v),\sigma(v_{1},\ldots,v_{\ell})=\sigma_{1}(v_{1})\vee\ldots\vee\sigma_{\ell}(v_{\ell}),

where σi:Rinu\sigma_{i}:R_{i}^{n}\rightarrow\mathbb{N}^{u} is a support defined via σi(vi):=σ(eiv)\sigma_{i}(v_{i}):=\sigma(e_{i}v), i[]i\in[\ell]. We claim that for each x[u]x\in[u] there is at most one i[]i\in[\ell] such that σi(vi)x0\sigma_{i}(v_{i})_{x}\neq 0 for some vRnv\in R^{n}. Indeed, assume towards a contradiction that there exist iji\neq j and v,wRnv,w\in R^{n} such that σi(vi)x,σj(wj)x0\sigma_{i}(v_{i})_{x},\sigma_{j}(w_{j})_{x}\neq 0. Without loss of generality we may assume that 0<σi(vi)xσj(wj)x0<\sigma_{i}(v_{i})_{x}\leq\sigma_{j}(w_{j})_{x}. By Property (P\star) there exists r=(r1,,r)Rr=(r_{1},\ldots,r_{\ell})\in R such that

σ(eiv)x>σ(eiv+ejrw)x=[σi(vi)σj(rjwj)]xσi(vi)x,\sigma(e_{i}v)_{x}>\sigma(e_{i}v+e_{j}rw)_{x}=[\sigma_{i}(v_{i})\vee\sigma_{j}(r_{j}w_{j})]_{x}\geq\sigma_{i}(v_{i})_{x},

where the equality follows from Proposition 2.13. This a contradiction, establishing the claim.

We have shown that for each x[u]x\in[u] there exists at most one i[]i\in[\ell] for which (σi)x(\sigma_{i})_{x} is not the zero function. In other words, the supports of the images of the functions σi\sigma_{i} are disjoint. Up to permuting the coordinates of u\mathbb{N}^{u}, one may assume that σ1\sigma_{1} is supported on the first u1u_{1} coordinates, σ2\sigma_{2} on the next u2u_{2},…, and σ\sigma_{\ell} on the last uu_{\ell}. Therefore one may regard each σi\sigma_{i} as a function which takes values in ui\mathbb{N}^{u_{i}}. Then σ=σ1××σ\sigma=\sigma_{1}\times\ldots\times\sigma_{\ell} and each σi\sigma_{i} is a modular support by Proposition 2.7(2) and Proposition 2.17(2). ∎

By combining Remark 2.14 with Theorem 2.23, support functions on a principal ideal ring RR can be easily characterized as follows.

Corollary 2.24.

Let RR be a finite principal ideal ring and let σ:Ru\sigma:R\rightarrow\mathbb{N}^{u} be a modular support. By the Zariski-Samuel Theorem, R=R1××RR=R_{1}\times\ldots\times R_{\ell} where R1,,RR_{1},\ldots,R_{\ell} are finite chain rings. For each ii, let αi\alpha_{i} be a generator of the maximal ideal of RiR_{i} and let ϵi:=min{j:αij=0}\epsilon_{i}:=\min\{j\,:\,\alpha_{i}^{j}=0\}. Then there exist u1,,uu_{1},\ldots,u_{\ell} such that u1++u=uu_{1}+\ldots+u_{\ell}=u and σ=σ1××σ\sigma=\sigma_{1}\times\ldots\times\sigma_{\ell}, where σi:Riui\sigma_{i}:R_{i}\rightarrow\mathbb{N}^{u_{i}} for i[]i\in[\ell]. Let σi(αij)=a(i,j)ui\sigma_{i}(\alpha_{i}^{j})=a^{(i,j)}\in\mathbb{N}^{u_{i}} for i[]i\in[\ell] and j{0,,ϵi1}j\in\{0,\ldots,\epsilon_{i}-1\}. Then (a(i,j1))k>(a(i,j))k(a^{(i,j-1)})_{k}>(a^{(i,j)})_{k} for j[ϵi1]j\in[\epsilon_{i}-1], i[]i\in[\ell], k[ui]k\in[u_{i}].

Conversely, any set of vectors a(i,j)uia^{(i,j)}\in\mathbb{N}^{u_{i}} such that a(i,j1)a(i,j)a^{(i,j-1)}\geq a^{(i,j)} for j[ϵi1]j\in[\epsilon_{i}-1] and i[]i\in[\ell] defines a support σ=σ1××σ\sigma=\sigma_{1}\times\ldots\times\sigma_{\ell} on RR via σi(rαij)=a(i,j)\sigma_{i}(r\alpha_{i}^{j})=a^{(i,j)} for i[],j{0,,ϵi1}i\in[\ell],j\in\{0,\ldots,\epsilon_{i}-1\}, and rRir\in R_{i} invertible. Moreover, if (a(i,j1))k>(a(i,j))k(a^{(i,j-1)})_{k}>(a^{(i,j)})_{k} for j[ϵi1]j\in[\epsilon_{i}-1], i[]i\in[\ell], and k[ui]k\in[u_{i}], then σ\sigma is modular.

The following is a reformulation of Property (P\star) for elements of 0:RnJ0:_{R^{n}}J.

Corollary 2.25.

Assume that σ\sigma is modular. If v,w0:RnJv,w\in 0:_{R^{n}}J and i[u]i\in[u] satisfy σ(v)i0\sigma(v)_{i}\neq 0 and σ(w)i0\sigma(w)_{i}\neq 0, then there exists a unit rRr\in R with σ(vrw)i=0\sigma(v-rw)_{i}=0.

Proof.

By Theorem 2.23 we may assume without loss of generality that RR is a finite local ring. Indeed, let j[]j\in[\ell] be such that (σj)i(\sigma_{j})_{i} is not identically zero and suppose that σj(vjrjwj)i=0\sigma_{j}(v_{j}-r_{j}w_{j})_{i}=0 for some rjRjr_{j}\in R_{j} invertible, then r=(1,,1,rj,1,,1)Rr=(1,\ldots,1,r_{j},1,\ldots,1)\in R is invertible and satisfies σ(vrw)i=0\sigma(v-rw)_{i}=0.

Assume now that (R,𝔐)(R,\mathfrak{M}) is a finite local ring. If σ(v)iσ(w)i\sigma(v)_{i}\leq\sigma(w)_{i}, then by Property (P\star) there is rRr\in R such that σ(vrw)i=0\sigma(v-rw)_{i}=0. If r𝔐r\in\mathfrak{M}, then rw=0rw=0, hence σ(v)i=0\sigma(v)_{i}=0, contradicting the assumption in the statement. Therefore rr is invertible. Similarly, if σ(w)iσ(v)i\sigma(w)_{i}\leq\sigma(v)_{i}, then there exists sRs\in R invertible such that 0=σ(wsv)i=σ(vs1w)i0=\sigma(w-sv)_{i}=\sigma(v-s^{-1}w)_{i}. ∎

3 Codewords and subcodes of minimal support

In this section we study the codewords and subcodes of minimal support of an RR-linear code endowed with a modular support. In particular, we establish some properties of the systems of generators of subcodes of minimal support. This allows us to derive properties of the generalized weights, such as monotonicity and a generalization of the Singleton bound.

In the sequel, we follow the notation of the previous sections and let σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} be a modular support. The minimal codewords of a code play a central role in our work. They are defined as follows.

Definition 3.1.

Let CRnC\subseteq R^{n} be a code. We say that vC{0}v\in C\setminus\{0\} is minimal in CC if its support is minimal among the supports of the elements of C{0}C\setminus\{0\}. We denote by Min(C)\textnormal{Min}(C) the set of minimal codewords of CC.

Remark 3.2.

By definition, C=0C=0 has no minimal codewords, i.e., Min(0)=\textnormal{Min}(0)=\emptyset.

We start by observing that a modular support σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} that takes values in {0,1}u\{0,1\}^{u} allows us to associate a matroid to a code CC. More precisely, the minimal ones among the supports of the codewords of CC are the circuits of a matroid. This generalizes the well-known fact that one may associate to a linear block-code the matroid represented by its parity-check matrix, whose circuits correspond to the minimal supports of the nonzero codewords of CC with respect to the Hamming weight.

Theorem 3.3.

Let RR be a finite commutative ring and let σ:Rn{0,1}u\sigma:R^{n}\to\{0,1\}^{u} be a modular support. Let 0CRn0\neq C\subseteq R^{n} be a code. Then the elements of the set

𝒞:={σ(v):vMin(C)}\mathcal{C}:=\{\sigma(v)\,:\,v\in\textnormal{Min}(C)\}

are the circuits of a matroid.

Proof.

If a support σ\sigma takes values in {0,1}u\{0,1\}^{u}, then the support of a vector can be naturally identified with a subset of [u][u]. In order to show that 𝒞\mathcal{C} is the set of circuits of a matroid, we check the circuit axioms as stated in [14, page 9].

Properties (C1) and (C2) are immediate to verify. To see that Property (C3) holds, suppose that σ(v),σ(w)𝒞\sigma(v),\sigma(w)\in\mathcal{C}, that σ(v)σ(w)\sigma(v)\neq\sigma(w), and that (σ(v)σ(w))i0(\sigma(v)\wedge\sigma(w))_{i}\neq 0. By repeatedly applying Property (P\star) and up to exchanging the role of vv and ww, one sees that there exists rRr\in R with σ(vrw)i=0\sigma(v-rw)_{i}=0. We claim that vrw0v-rw\neq 0. Indeed, if v=rwv=rw then we would have σ(v)=σ(rw)σ(w)\sigma(v)=\sigma(rw)\leq\sigma(w). Since σ(w)\sigma(w) is minimal by assumption and v0v\neq 0, it must be that σ(v)=σ(w)\sigma(v)=\sigma(w), a contradiction.

Since vrw0v-rw\neq 0, we have σ(vrw)0\sigma(v-rw)\neq 0. Fix zCz\in C with σ(z)𝒞\sigma(z)\in\mathcal{C}, σ(z)σ(vrw)\sigma(z)\leq\sigma(v-rw). We have

σ(z)σ(vrw)σ(v)σ(rw)σ(v)σ(w).\sigma(z)\leq\sigma(v-rw)\leq\sigma(v)\vee\sigma(-rw)\leq\sigma(v)\vee\sigma(w). (3.1)

Moreover, 0=σ(vrw)iσ(z)i0=\sigma(v-rw)_{i}\geq\sigma(z)_{i}. This establishes Property (C3). ∎

We start our study of the minimal codewords by showing that the minimal codewords of a code CC coincide with those of its socle. We also show that the minimal codewords are determined by their support, up to multiplication by a unit.

Theorem 3.4.

Let CRnC\subseteq R^{n} be a code and assume that σ\sigma is modular. The following hold.

  1. (1)

    The set of minimal codewords of CC is

    Min(C)\displaystyle\textnormal{Min}(C) =\displaystyle= i=1(0××0×Min(Ci)×0××0)\displaystyle\bigcup_{i=1}^{\ell}(0\times\ldots\times 0\times\textnormal{Min}(C_{i})\times 0\times\ldots\times 0)
    \displaystyle\subseteq i=1(0××0×(0:Ci𝔐i)×0××0)  0:CJ.\displaystyle\bigcup_{i=1}^{\ell}(0\times\ldots\times 0\times(0:_{C_{i}}\mathfrak{M}_{i})\times 0\times\ldots\times 0)\;\;\subseteq\;\;0:_{C}J.
  2. (2)

    In particular,

    Min(C)=Min(0:CJ).\textnormal{Min}(C)=\textnormal{Min}(0:_{C}J).
  3. (3)

    If v,wMin(C)v,w\in\textnormal{Min}(C) are minimal codewords with σ(v)=σ(w)\sigma(v)=\sigma(w), then v=rwv=rw for some invertible rRr\in R.

Proof.
  1. (1)

    By Theorem 2.23, up to a permutation of the coordinates of u\mathbb{N}^{u}, σ\sigma decomposes as a product σ=σ1××σ\sigma=\sigma_{1}\times\ldots\times\sigma_{\ell}, where each σi\sigma_{i} is a modular support. Let i[]i\in[\ell] and v=(v1,,v)Min(C)v=(v_{1},\ldots,v_{\ell})\in\textnormal{Min}(C) with vi0v_{i}\neq 0. Then 0σ(eiv)σ(v)0\neq\sigma(e_{i}v)\leq\sigma(v), hence σ(eiv)=σ(v)\sigma(e_{i}v)=\sigma(v). In particular, σj(vj)=0\sigma_{j}(v_{j})=0 for all jij\neq i, hence vj=0v_{j}=0 for all jij\neq i. Therefore, v=eivv=e_{i}v and viMin(Ci)v_{i}\in\textnormal{Min}(C_{i}). This proves the equality in the statement.

    Suppose now that (R,𝔐)(R,\mathfrak{M}) is a finite local ring and vMin(C)v\in\textnormal{Min}(C). If rRr\in R is such that rv0rv\neq 0, then σ(v)=σ(rv)\sigma(v)=\sigma(rv). Since σ\sigma is modular, there exists sRs\in R such that σ(vsrv)<σ(v)\sigma(v-srv)<\sigma(v), hence vsrv=0v-srv=0 by the minimality of σ(v)\sigma(v). Hence 1sr0:Rv𝔐1-sr\in 0:_{R}v\subseteq\mathfrak{M}. This shows that sr𝔐sr\not\in\mathfrak{M}, hence r𝔐r\not\in\mathfrak{M}. Therefore, v0:C𝔐v\in 0:_{C}\mathfrak{M}, which proves the second inclusion.

    The inclusion follows from the fact that 0:CJ=(0:C1𝔐1)××(0:C𝔐)0:_{C}J=(0:_{C_{1}}\mathfrak{M}_{1})\times\ldots\times(0:_{C_{\ell}}\mathfrak{M}_{\ell}).

  2. (2)

    This follows from part (1), since 0:CJC0:_{C}J\subseteq C implies

    Min(0:CJ)Min(C)(0:CJ)=Min(C),\textnormal{Min}(0:_{C}J)\supseteq\textnormal{Min}(C)\cap(0:_{C}J)=\textnormal{Min}(C),

    where the equality follows from part (1). Conversely, if vMin(0:CJ)v\in\textnormal{Min}(0:_{C}J), then there exists wMin(C)w\in\textnormal{Min}(C) such that σ(w)σ(v)\sigma(w)\leq\sigma(v). Since w0:CJw\in 0:_{C}J by part (1), then σ(w)=σ(v)\sigma(w)=\sigma(v) and vMin(C)v\in\textnormal{Min}(C).

  3. (3)

    Since σ\sigma is modular, there exists rRr\in R such that σ(vrw)<σ(v)\sigma(v-rw)<\sigma(v). By the minimality of σ(v)\sigma(v), vrw=0v-rw=0, hence v=rwv=rw. Exchanging the roles of vv and ww one sees that there exists sRs\in R such that w=svw=sv. Therefore, (1rs)v=0(1-rs)v=0, so 1rs0:Rv1-rs\in 0:_{R}v. By part (1), v=eivv=e_{i}v for some i[]i\in[\ell] and 0:Rv=R1××Ri1×𝔐i×Ri+1××R0:_{R}v=R_{1}\times\ldots\times R_{i-1}\times\mathfrak{M}_{i}\times R_{i+1}\times\ldots\times R_{\ell}. Since 1siri𝔐i1-s_{i}r_{i}\in\mathfrak{M}_{i}, then ri𝔐ir_{i}\not\in\mathfrak{M}_{i}, hence rir_{i} is invertible. Let r¯=(1,,1,ri,1,,1)\overline{r}=(1,\ldots,1,r_{i},1,\ldots,1). Then v=r¯wv=\overline{r}w and r¯R\overline{r}\in R is invertible. ∎

Theorem 3.4 implies that a code generated by its minimal codewords must be a subcode of 0:RnJ0:_{R^{n}}J. In the next theorem we prove that every subcode of 0:RnJ0:_{R^{n}}J is generated by its minimal codewords.

Theorem 3.5.

Let 0CRn0\neq C\subseteq R^{n} be a code and assume that σ\sigma is modular. Then 0:CJ0:_{C}J has a minimal system of generators consisting of codewords that are minimal in CC. Moreover, every minimal system of generators of 0:CJ0:_{C}J consisting of minimal codewords has the same cardinality M(0:CJ)M(0:_{C}J). In particular, CC has a minimal system of generators consisting of minimal codewords if and only if C0:RnJC\subseteq 0:_{R^{n}}J. If this is the case, then every such minimal system of generators has cardinality M(C)M(C).

Proof.

By Theorem 3.4(2) we have Min(C)=Min(0:CJ)\textnormal{Min}(C)=\textnormal{Min}(0:_{C}J). Let D=0:CJD=0:_{C}J. Since every system of generators of DD contains a minimal one, in order to show that DD has a minimal system of generators consisting of minimal codewords, it suffices to show that the elements of Min(D)\textnormal{Min}(D) generate DD.

Let vDv\in D and suppose by contradiction that vv is a codeword of minimal support among those in the set DMin(D)D\setminus\langle\textnormal{Min}(D)\rangle. Since vMin(D)v\not\in\textnormal{Min}(D), then there is a wMin(D)w\in\textnormal{Min}(D) such that σ(w)<σ(v)\sigma(w)<\sigma(v). Let i[u]i\in[u] such that σ(w)i0\sigma(w)_{i}\neq 0. By Theorem 3.4(1), w=ejww=e_{j}w for some j[]j\in[\ell] and wj𝔐j=0w_{j}\mathfrak{M}_{j}=0. By Property (P\star) there exists rRr\in R such that σ(wrv)i=0\sigma(w-rv)_{i}=0. Since

0=σ(wrv)i=σ(ejwejrv)i<σ(ejw)i=σ(w)i,0=\sigma(w-rv)_{i}=\sigma(e_{j}w-e_{j}rv)_{i}<\sigma(e_{j}w)_{i}=\sigma(w)_{i}, (3.2)

then rj𝔐jr_{j}\not\in\mathfrak{M}_{j}. Indeed, v0:CJv\in 0:_{C}J implies that vj0:Cj𝔐jv_{j}\in 0:_{C_{j}}\mathfrak{M}_{j}. Hence, if rj𝔐jr_{j}\in\mathfrak{M}_{j}, then ejwejrv=ejwe_{j}w-e_{j}rv=e_{j}w, contradictiong equation (3.2). Let s=(1,,1,rj,1,,1)Rs=(1,\ldots,1,r_{j},1,\ldots,1)\in R. Then sRs\in R is invertible and

σ(wsv)i=σ(ejwejsv)i=σ(ejwejrv)i=σ(wrv)i=0.\sigma(w-sv)_{i}=\sigma(e_{j}w-e_{j}sv)_{i}=\sigma(e_{j}w-e_{j}rv)_{i}=\sigma(w-rv)_{i}=0.

Then σ(wsv)<σ(v)\sigma(w-sv)<\sigma(v), hence by the minimality of σ(v)\sigma(v) among the supports of elements of DMin(D)D\setminus\langle\textnormal{Min}(D)\rangle we have that wsvMin(D)w-sv\in\langle\textnormal{Min}(D)\rangle. Since ss is invertible, this implies that vMin(D)v\in\langle\textnormal{Min}(D)\rangle, which contradicts the assumption that vDMin(D)v\in D\setminus\langle\textnormal{Min}(D)\rangle.

In order to prove that every minimal system of generators of DD consisting of minimal codewords has cardinality M(D)M(D), write D=D1××DD=D_{1}\times\ldots\times D_{\ell}. By Theorem 3.4(1), every minimal codeword vv of DD satisfies v=eivv=e_{i}v for some i[]i\in[\ell]. Therefore, each minimal system of generators of DD consisting of minimal codewords is the union for i[]i\in[\ell] of minimal systems of generators of 0××0×Di×0××00\times\ldots\times 0\times D_{i}\times 0\times\ldots\times 0. Since RiR_{i} is a finite local ring, the cardinality of any minimal system of generators of 0××0×Di×0××00\times\ldots\times 0\times D_{i}\times 0\times\ldots\times 0 is μ(Di)\mu(D_{i}) by Theorem 1.5. Therefore, the cardinality of a minimal system of generators of DD consisting of minimal codewords is μ(D1)=+μ(D)=M(D)\mu(D_{1})=\ldots+\mu(D_{\ell})=M(D). ∎

We stress that not every minimal system of generators of a code C0:RnJC\subseteq 0:_{R^{n}}J consists of minimal codewords.

Example 3.6.

The element (2,3)(2,3) is not an element of minimal support in C=(2,3)62C=\langle(2,3)\rangle\subseteq\mathbb{Z}_{6}^{2}. However, (2,0),(0,3)(2,0),(0,3) are elements of minimal support that generate CC. Here C1=(2,0)32C_{1}=\langle(2,0)\rangle\subseteq\mathbb{Z}_{3}^{2}, C2=(0,1)22C_{2}=\langle(0,1)\rangle\subseteq\mathbb{Z}_{2}^{2}, and M(C)=μ(C1)+μ(C2)=2M(C)=\mu(C_{1})+\mu(C_{2})=2.

The following property of minimal codewords will be needed in the next proposition.

Lemma 3.7.

Let σ\sigma be a modular support on RnR^{n}. Let C0:RnJC\subseteq 0:_{R^{n}}J be a code and let vCv\in C with σ(v)i0\sigma(v)_{i}\neq 0. Then there exists wMin(C)w\in\textnormal{Min}(C) with σ(w)σ(v)\sigma(w)\leq\sigma(v) and σ(w)i0\sigma(w)_{i}\neq 0.

Proof.

Write C=C1××CC=C_{1}\times\ldots\times C_{\ell}. By Theorem 2.23, we may assume without loss of generality that (R,𝔐)(R,\mathfrak{M}) is a finite local ring. Indeed, if σ=σ1×σ\sigma=\sigma_{1}\times\ldots\sigma_{\ell} and σ(v)i=σk(vk)i\sigma(v)_{i}=\sigma_{k}(v_{k})_{i} for some k[]k\in[\ell], then ekv0××0×Rk×0××0e_{k}v\in 0\times\ldots\times 0\times R_{k}\times 0\times\ldots\times 0 has σ(ekv)σ(v)\sigma(e_{k}v)\leq\sigma(v) and σ(ekv)i0\sigma(e_{k}v)_{i}\neq 0. If wkMin(Ck)w_{k}\in\textnormal{Min}(C_{k}) and σk(wk)i0\sigma_{k}(w_{k})_{i}\neq 0, then σk(wk)σk(vk)\sigma_{k}(w_{k})\leq\sigma_{k}(v_{k}), therefore

σ(0,,0,wk,0,,0)σ(ekv)σ(v)\sigma(0,\ldots,0,w_{k},0,\ldots,0)\leq\sigma(e_{k}v)\leq\sigma(v)

and σ(0,,0,wk,0,,0)i=σk(wk)i0\sigma(0,\ldots,0,w_{k},0,\ldots,0)_{i}=\sigma_{k}(w_{k})_{i}\neq 0. Moreover, (0,,0,wk,0,,0)Min(C)(0,\ldots,0,w_{k},0,\ldots,0)\in\textnormal{Min}(C) as wkMin(Ck)w_{k}\in\textnormal{Min}(C_{k}).

Proceed by induction on |σ(v)||\sigma(v)|. Let vMin(C)v^{\prime}\in\textnormal{Min}(C) with σ(v)σ(v)\sigma(v^{\prime})\leq\sigma(v). If σ(v)i0\sigma(v^{\prime})_{i}\neq 0 then let w=vw=v^{\prime}, else fix j[u]j\in[u] with σ(v)j0\sigma(v^{\prime})_{j}\neq 0. By Corollary 2.25 there exists rRr\in R invertible such that σ(vrv)j=0\sigma(v^{\prime}-rv)_{j}=0. Hence σ(vrv)<σ(v)\sigma(v^{\prime}-rv)<\sigma(v) and σ(vrv)i0\sigma(v^{\prime}-rv)_{i}\neq 0, by Lemma 2.3(P5). So we may apply the induction hypothesis to vrvv^{\prime}-rv and obtain wMin(C)w\in\textnormal{Min}(C) such that σ(w)σ(vrv)σ(v)\sigma(w)\leq\sigma(v^{\prime}-rv)\leq\sigma(v) and σ(w)i0\sigma(w)_{i}\neq 0. ∎

We now prove that modularity allows us to produce minimal system of generators of submodules of 0:RnJ0:_{R^{n}}J, whose supports have a shape which is reminiscent of the rows of a matrix in reduced row-echelon form.

Proposition 3.8.

Let j1j\geq 1 and let C𝒮j(0:RnJ)C\in\mathcal{S}_{j}(0:_{R^{n}}J). If σ\sigma is modular, then CC has a minimal system of generators {v1,,vj}\{v_{1},\ldots,v_{j}\} such that for all i[j]i\in[j] there exists ki[u]k_{i}\in[u] with σ(vi)ki0\sigma(v_{i})_{k_{i}}\neq 0 and σ(vh)ki=0\sigma(v_{h})_{k_{i}}=0 for all hih\neq i.

Proof.

Any system of generators with the required property is minimal, since for all i[j]i\in[j]

σ(vi)hiσ(vh).\sigma(v_{i})\not\leq\bigvee_{h\neq i}\sigma(v_{h}).

We prove that CC has such a system of generators by induction on jj. The statement is trivial if j=1j=1. Hence assume j2j\geq 2 and fix a minimal system of generators {w1,,wj}\{w_{1},\ldots,w_{j}\} of CC. Up to permuting the entries in the supports, we may assume without loss of generality that σ(w1)10\sigma(w_{1})_{1}\neq 0. By Corollary 2.25 there exist r2,,rjRr_{2},\ldots,r_{j}\in R with σ(wiriw1)1=0\sigma(w_{i}-r_{i}w_{1})_{1}=0 for all i{2,,j}i\in\{2,\ldots,j\}. Let wi:=wiriw1w^{\prime}_{i}:=w_{i}-r_{i}w_{1} for i{2,,j}i\in\{2,\ldots,j\} and observe that CC is generated by {w1,w2,,wj}\{w_{1},w_{2}^{\prime},\ldots,w_{j}^{\prime}\}. Moreover, σ(v)1=0\sigma(v)_{1}=0 for all vC=w2,,wjv\in C^{\prime}=\langle w_{2}^{\prime},\ldots,w_{j}^{\prime}\rangle by Lemma 2.3(P6). We apply the induction hypothesis to the code C=w2,,wjC^{\prime}=\langle w_{2}^{\prime},\ldots,w_{j}^{\prime}\rangle, obtaining a system of generators {v2,,vj}\{v_{2},\ldots,v_{j}\} of CC^{\prime} such that for all i{2,,j}i\in\{2,\ldots,j\} there exists kik_{i} with σ(vi)ki0\sigma(v_{i})_{k_{i}}\neq 0 and σ(vh)ki=0\sigma(v_{h})_{k_{i}}=0 for h{2,,j}{i}h\in\{2,\ldots,j\}\setminus\{i\}. By Corollary 2.25 we find r2,,rjRr_{2}^{\prime},\ldots,r_{j}^{\prime}\in R with σ(w1rivi)ki=0\sigma(w_{1}-r_{i}^{\prime}v_{i})_{k_{i}}=0 for i{2,,j}i\in\{2,\ldots,j\}. Finally, let v1:=w1i=2jriviv_{1}:=w_{1}-\sum_{i=2}^{j}r_{i}^{\prime}v_{i} and set k1=1k_{1}=1. By parts (P5) and (P6) of Lemma 2.3 we have σ(v1)k10\sigma(v_{1})_{k_{1}}\neq 0 and σ(v1)ki=0\sigma(v_{1})_{k_{i}}=0 for i{2,,j}i\in\{2,\ldots,j\}. In addition, {v1,,vj}\{v_{1},\ldots,v_{j}\} is a system of generators of CC, since {w1,v2,,vj}\{w_{1},v_{2},\ldots,v_{j}\} is. ∎

The proposition also implies that the codomain of a modular support cannot be too small.

Corollary 3.9.

If σ:Rnu\sigma:R^{n}\to\mathbb{N}^{u} is modular, then uM(0:RnJ)u\geq M(0:_{R^{n}}J). In particular, if RR is a PIR or JRn=0JR^{n}=0, then unu\geq\ell n.

Proof.

Proposition 3.8 for C=0:RnJC=0:_{R^{n}}J and j=M(0:RnJ)j=M(0:_{R^{n}}J) implies that uM(0:RnJ)u\geq M(0:_{R^{n}}J). If JRn=0JR^{n}=0, then Rn=0:RnJR^{n}=0:_{R^{n}}J, hence M(0:RnJ)=M(Rn)M(0:_{R^{n}}J)=M(R^{n}). If RR is a PIR, then M(0:RnJ)=M(Rn)M(0:_{R^{n}}J)=M(R^{n}) by Proposition 1.10. In both cases, one has that M(0:RnJ)=nM(0:_{R^{n}}J)=\ell n. ∎

Understanding the subcodes of CC generated by minimal codewords allows us to prove that the generalized of weights of CC are attained by subcodes of 0:CJ0:_{C}J. In particular, CC and its socle have the same generalized weights.

Proposition 3.10.

Let RR be a PIR. Suppose that σ\sigma is modular and let CRnC\subseteq R^{n} be a code. Let r[M(C)]r\in[M(C)] and D𝒮j(C)D\in\mathcal{S}_{j}(C), jrj\geq r, be such that dr(C)=|σ(D)|d_{r}(C)=|\sigma(D)|. Then 0:DJ𝒮i(C)0:_{D}J\in\mathcal{S}_{i}(C) for some iri\geq r and dr(C)=|σ(0:DJ)|.d_{r}(C)=|\sigma(0:_{D}J)|. In particular,

dr(C)=dr(0:CJ).d_{r}(C)=d_{r}(0:_{C}J).
Proof.

By Proposition 1.10 and Theorem 3.5, 0:DJD0:_{D}J\subseteq D is minimally generated by a set of M(0:DJ)=M(D)M(0:_{D}J)=M(D) codewords. Therefore 0:DJ𝒮M(D)(C)0:_{D}J\in\mathcal{S}_{M(D)}(C) and M(D)rM(D)\geq r. Moreover

|σ(0:DJ)||σ(D)|=dr(C),|\sigma(0:_{D}J)|\leq|\sigma(D)|=d_{r}(C),

hence equality holds. Since 0:DJ0:CJ0:_{D}J\subseteq 0:_{C}J, then

dr(0:CJ)|σ(0:DJ)|=dr(C).d_{r}(0:_{C}J)\leq|\sigma(0:_{D}J)|=d_{r}(C).

The reverse inequality follows from the inclusion 0:CJC0:_{C}J\subseteq C. ∎

In particular, this allows us to determine the last generalized weight of CC.

Corollary 3.11.

Let CRnC\subseteq R^{n} be a code and σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Then

dM(C)(C)=|σ(0:CJ)|.d_{M(C)}(C)=|\sigma(0:_{C}J)|.
Proof.

We claim that M(C)=M(0:CJ)M(C)=M(0:_{C}J) and dM(C)(C)=dM(C)(0:CJ)d_{M(C)}(C)=d_{M(C)}(0:_{C}J). This is clear if C0:RnJC\subseteq 0:_{R^{n}}J, since C=0:CJC=0:_{C}J. If RR is a PIR, the claim follows from Proposition 1.10 and Proposition 3.10.

By Proposition 1.9, M(D)M(C)M(D)\leq M(C) for every D0:CJD\subseteq 0:_{C}J and the only subcode D0:CJD\subseteq 0:_{C}J with M(D)=M(C)M(D)=M(C) is D=0:CJD=0:_{C}J. Therefore

dM(C)(C)=|σ(0:CJ)|.d_{M(C)}(C)=|\sigma(0:_{C}J)|.\qed

For a given code, we can produce subcodes that attain its generalized weights and that are minimally generated by a set of minimal codewords, whose supports have the same reduced shape as in Proposition 3.8. This technical result plays a crucial role in the proof of Theorem 4.4.

Theorem 3.12.

Let CRnC\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Then, for all r[M(C)]r\in[M(C)], there exists a subcode DCD\subseteq C such that:

  1. 1.

    dr(C)=|σ(D)|d_{r}(C)=|\sigma(D)|,

  2. 2.

    DD has a minimal system of generators {v1,,vr}\{v_{1},\ldots,v_{r}\} such that viMin(C)v_{i}\in\textnormal{Min}(C) for all i[r]i\in[r]. Moreover

    σ(vi)jiσ(vj).\sigma(v_{i})\not\leq\bigvee_{j\neq i}\sigma(v_{j}).
Proof.

If C0:RnJC\subseteq 0:_{R^{n}}J, then let DCD^{\prime}\subseteq C such that D𝒮j(C)D^{\prime}\in\mathcal{S}_{j}(C), jrj\geq r, and dr(C)=|σ(D)|d_{r}(C)=|\sigma(D^{\prime})|. If RR is a PIR, then by Proposition 3.10 there exist jrj\geq r and D0:CJD^{\prime}\subseteq 0:_{C}J such that D𝒮j(C)D^{\prime}\in\mathcal{S}_{j}(C) and dr(C)=|σ(D)|d_{r}(C)=|\sigma(D^{\prime})|. In both cases, by Proposition 3.8, DD^{\prime} has a minimal system of generators w1,,wjw_{1},\ldots,w_{j} with the following property: For all i[j]i\in[j] there exists ki[u]k_{i}\in[u] with σ(wi)ki0\sigma(w_{i})_{k_{i}}\neq 0 and σ(wh)ki=0\sigma(w_{h})_{k_{i}}=0 for hih\neq i. By Lemma 3.7, for all i[j]i\in[j] there exists viMin(C)v_{i}\in\textnormal{Min}(C) with σ(vi)σ(wi)\sigma(v_{i})\leq\sigma(w_{i}) and σ(vi)ki0\sigma(v_{i})_{k_{i}}\neq 0. In particular, σ(vi)hiσ(vh)\sigma(v_{i})\not\leq\vee_{h\neq i}\sigma(v_{h}) for i[j]i\in[j]. Let D=v1,,vrD=\langle v_{1},\ldots,v_{r}\rangle. Notice that D𝒮r(C)D\in\mathcal{S}_{r}(C), since vhvk:k[r],khv_{h}\not\in\langle v_{k}\,:\,k\in[r],k\neq h\rangle for all h[r]h\in[r]. Moreover,

|σ(D)|=i=1rσ(vi)i=1jσ(wi)=|σ(D)|.|\sigma(D)|=\bigvee_{i=1}^{r}\sigma(v_{i})\leq\bigvee_{i=1}^{j}\sigma(w_{i})=|\sigma(D^{\prime})|.

Therefore |σ(D)|=|σ(D)|=dr(C)|\sigma(D)|=|\sigma(D^{\prime})|=d_{r}(C). ∎

Notation 3.13.

Let CRnC\subseteq R^{n} be a code and let j[M(C)]j\in[M(C)]. We let

j(C):={DC:D has a minimal system of generators of j minimal codewords}.\mathcal{M}_{j}(C):=\{D\subseteq C\,:\,\mbox{$D$ has a minimal system of generators of $j$ minimal codewords}\}.

Theorem 3.12 shows that for all r[M(C)]r\in[M(C)] there exists Dr(C)D\in\mathcal{M}_{r}(C) such that dr(C)=|σ(D)|d_{r}(C)=|\sigma(D)|. In particular, we have shown the following.

Corollary 3.14.

Let CRnC\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. The following quantities are equal to the rr-th generalized weight dr(C)d_{r}(C), for any r[M(C)]r\in[M(C)]:

  1. 1.

    min{|σ(D)|:D𝒮j(C) for some jr}\min\{|\sigma(D)|\,:\,D\in\mathcal{S}_{j}(C)\mbox{ for some }j\geq r\},

  2. 2.

    min{|σ(D)|:D𝒮r(C)}\min\{|\sigma(D)|\,:\,D\in\mathcal{S}_{r}(C)\},

  3. 3.

    min{|σ(D)|:Dj(C) for some jr}\min\{|\sigma(D)|\,:\,D\in\mathcal{M}_{j}(C)\mbox{ for some }j\geq r\},

  4. 4.

    min{|σ(D)|:Dr(C)}.\min\{|\sigma(D)|\,:\,D\in\mathcal{M}_{r}(C)\}.

Proof.

Equality between dr(C)d_{r}(C) and 1. holds by definition. Equality between dr(C)d_{r}(C) and 4. follows directly from Theorem 3.12. Equality between dr(C)d_{r}(C) and 3. then follows from the chain of inclusions

r(C)jrj(C)jr𝒮j(C).\mathcal{M}_{r}(C)\subseteq\cup_{j\geq r}\mathcal{M}_{j}(C)\subseteq\cup_{j\geq r}\mathcal{S}_{j}(C).

Similarly, equality between dr(C)d_{r}(C) and 2. follows from the chain of inclusions

r(C)𝒮r(C)jr𝒮j(C).\mathcal{M}_{r}(C)\subseteq\mathcal{S}_{r}(C)\subseteq\cup_{j\geq r}\mathcal{S}_{j}(C).\qed
Remark 3.15.

Theorem 3.5 and Corollary 3.14 are in general false for supports that are not modular. For instance, the support of Example 2.21 violates both results taking C=𝔽22C=\mathbb{F}_{2}^{2}.

We can now prove that the generalized weights form a strictly increasing sequence. This extends a classical result by Wei [17].

Theorem 3.16.

Let CRnC\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Then

minwt(C)=d1(C)<d2(C)<<dM(C)(C)=|σ(0:CJ)|.\min\textnormal{wt}(C)=d_{1}(C)<d_{2}(C)<\ldots<d_{M(C)}(C)=|\sigma(0:_{C}J)|.
Proof.

By Corollary 3.14 and Theorem 3.4(1) we may assume without loss of generality that C0:RnJC\subseteq 0:_{R^{n}}J. Let r[M(C)1]r\in[M(C)-1] and let DCD\subseteq C be such that |σ(D)|=dr+1(C)|\sigma(D)|=d_{r+1}(C). We may assume that DD has a minimal system of generators {v1,,vj+1}\{v_{1},\ldots,v_{j+1}\} as in Proposition 3.8 with jrj\geq r. Then D:=v1,,vj𝒮j(D)D^{\prime}:=\langle v_{1},\ldots,v_{j}\rangle\in\mathcal{S}_{j}(D). We have σ(D)σ(D)\sigma(D^{\prime})\leq\sigma(D) and σ(D)kj+1=0<σ(D)kj+1\sigma(D^{\prime})_{k_{j+1}}=0<\sigma(D)_{k_{j+1}}, hence |σ(D)|<|σ(D)||\sigma(D^{\prime})|<|\sigma(D)|. In particular,

dr(C)|σ(D)|<|σ(D)|=dr+1(C).d_{r}(C)\leq|\sigma(D^{\prime})|<|\sigma(D)|=d_{r+1}(C).

The two equalities in the statement follow from Lemma 2.12(1) and Corollary 3.11. ∎

As an application of Theorem 3.16, we extend the generalized Singleton bound [17, Corollary 1] to every code over a PIR and some codes over finite commutative rings.

Corollary 3.17.

Let CRnC\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Then

minwt(C)+r1dr(C)|σ(0:RnJ)|M(C)+r\min\textnormal{wt}(C)+r-1\leq d_{r}(C)\leq|\sigma(0:_{R^{n}}J)|-M(C)+r

for all r[M(C)]r\in[M(C)]. In particular,

minwt(C)|σ(0:RnJ)|M(C)+1.\min\textnormal{wt}(C)\leq|\sigma(0:_{R^{n}}J)|-M(C)+1.
Proof.

The result follows by combining Theorem 3.16 with

dM(C)=|σ(0:CJ)||σ(0:RnJ)|,d_{M}(C)=|\sigma(0:_{C}J)|\leq|\sigma(0:_{R^{n}}J)|,

where the equality on the left hand side follows from Corollary 3.11. ∎

The next corollary proves that, for modular supports, any subcode DD of CC with dr(C)=|σ(D)|d_{r}(C)=|\sigma(D)| has a minimal system of generators consisting of rr elements, and no minimal system of generators of larger cardinality. This result allows us to restrict to such subcodes when studying the generalized weights of CC.

Corollary 3.18.

Let CRnC\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Let r[M(C)]r\in[M(C)] and D𝒮j(C)D\in\mathcal{S}_{j}(C), jrj\geq r, be such that dr(C)=|σ(D)|d_{r}(C)=|\sigma(D)|. Then r=j=M(D)r=j=M(D).

Proof.

Since D𝒮j(C)D\in\mathcal{S}_{j}(C), then rjM(D)r\leq j\leq M(D) and D𝒮M(D)(C)D\in\mathcal{S}_{M(D)}(C) by Proposition 1.8. Then

|σ(D)|dM(D)(C)dr(C)=|σ(D)|,|\sigma(D)|\geq d_{M(D)}(C)\geq d_{r}(C)=|\sigma(D)|,

where the first inequality follows from D𝒮M(D)(C)D\in\mathcal{S}_{M(D)}(C) and the second from Lemma 2.12(3). Therefore the inequalities are equalities and r=j=M(D)r=j=M(D) by Theorem 3.16. ∎

In the next theorem, we establish some additional properties of the subcodes of CC that realize the generalized weights of CC.

Theorem 3.19.

Let CRnC\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Let r[M(C)]r\in[M(C)] and D𝒮r(C)D\in\mathcal{S}_{r}(C) be such that dr(C)=|σ(D)|d_{r}(C)=|\sigma(D)|. The following hold.

  1. (1)

    If vMin(C)v\in\textnormal{Min}(C) satisfies σ(v)σ(D)\sigma(v)\leq\sigma(D), then vDv\in D. In particular, Min(D)=Min(C)D\textnormal{Min}(D)=\textnormal{Min}(C)\cap D.

  2. (2)

    0:DJ=vC:vMin(C),σ(v)σ(D)0:_{D}J=\langle v\in C\,:\,v\in\textnormal{Min}(C),\,\sigma(v)\leq\sigma(D)\rangle.

In particular, if Dr(C)D\in\mathcal{M}_{r}(C), then D=vC:vMin(C),σ(v)σ(D)D=\langle v\in C\,:\,v\in\textnormal{Min}(C),\,\sigma(v)\leq\sigma(D)\rangle.

Proof.
  1. (1)

    If C0:RnJC\subseteq 0:_{R^{n}}J, then also D0:RnJD\subseteq 0:_{R^{n}}J. If RR is a PIR, then 0:DJD0:_{D}J\subseteq D has σ(0:DJ)=σ(D)\sigma(0:_{D}J)=\sigma(D) by Proposition 3.10. In both cases, it suffices to prove the thesis under the assumption that D0:RnJD\subseteq 0:_{R^{n}}J.

    If vDv\not\in D, consider DD=D+v0:RnJD\subsetneq D^{\prime}=D+\langle v\rangle\subseteq 0:_{R^{n}}J. We have

    r=M(D)<M(D)M(D)+1=r+1,r=M(D)<M(D^{\prime})\leq M(D)+1=r+1,

    where the equalities follows from Corollary 3.18 and the first inequality follows from Proposition 1.9. The second inequality follows from observing that, if D=D1××DD=D_{1}\times\ldots\times D_{\ell} and v=ejvv=e_{j}v, then

    D=D1××Dj1×(Dj+vj)×Dj+1××D.D^{\prime}=D_{1}\times\ldots\times D_{j-1}\times(D_{j}+\langle v_{j}\rangle)\times D_{j+1}\times\ldots\times D_{\ell}.

    Since vjDjv_{j}\not\in D_{j} and 0:Rj𝔐jDj{vj}0:_{R_{j}}\mathfrak{M}_{j}\supseteq D_{j}\cup\{v_{j}\}, then

    μ(Dj+vj)=dimRj/𝔐j(Dj+vj)=dimRj/𝔐j(Dj)+1=μ(Dj)+1.\mu(D_{j}+\langle v_{j}\rangle)=\dim_{R_{j}/\mathfrak{M}_{j}}(D_{j}+\langle v_{j}\rangle)=\dim_{R_{j}/\mathfrak{M}_{j}}(D_{j})+1=\mu(D_{j})+1.

    Therefore M(D)=r+1M(D^{\prime})=r+1. Since σ(D)=σ(v)σ(D)=σ(D)\sigma(D^{\prime})=\sigma(v)\vee\sigma(D)=\sigma(D), then dr+1(C)|σ(D)|=dr(C)d_{r+1}(C)\leq|\sigma(D)|=d_{r}(C), contradicting Theorem 3.16. It follows that vDv\in D, as desired.

    In order to prove that Min(D)=Min(C)D\textnormal{Min}(D)=\textnormal{Min}(C)\cap D, it suffices to prove that Min(D)Min(C)\textnormal{Min}(D)\subseteq\textnormal{Min}(C). Let wMin(D)Cw\in\textnormal{Min}(D)\subseteq C, then there is a vMin(C)v\in\textnormal{Min}(C) such that σ(v)σ(w)σ(D)\sigma(v)\leq\sigma(w)\leq\sigma(D), where the second inequality follows from wDw\in D. By the first part of the proof, vDv\in D. Therefore σ(w)=σ(v)\sigma(w)=\sigma(v) and wMin(C)w\in\textnormal{Min}(C).

  2. (2)

    By Theorem 3.4(2) and Theorem 3.5 and part (1),

    0:DJ=vMin(D)=vMin(C)D=vMin(C):σ(v)σ(D).0:_{D}J=\langle v\in\textnormal{Min}(D)\rangle=\langle v\in\textnormal{Min}(C)\cap D\rangle=\langle v\in\textnormal{Min}(C)\,:\,\sigma(v)\leq\sigma(D)\rangle.\qed

The next result relates the generalized weights of CC with those of its factors.

Corollary 3.20.

Let C=C1××CRnC=C_{1}\times\ldots\times C_{\ell}\subseteq R^{n} be a code and let σ\sigma be a modular support. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Then

dr(C)=min{j=1drj(Cj):r1++r=r,rj{0,,μ(Cj)}}d_{r}(C)=\min\left\{\sum_{j=1}^{\ell}d_{r_{j}}(C_{j})\,:\,r_{1}+\ldots+r_{\ell}=r,\;r_{j}\in\{0,\ldots,\mu(C_{j})\}\right\}

for all r[M(C)]r\in[M(C)].

Proof.

By Theorem 2.23, up to a permutation of the coordinates of u\mathbb{N}^{u} we can write σ=σ1××σ\sigma=\sigma_{1}\times\ldots\times\sigma_{\ell}, where σj:Rjnuj\sigma_{j}:R_{j}^{n}\to\mathbb{N}^{u_{j}} is a modular support for j[]j\in[\ell] and u=u1++uu=u_{1}+\ldots+u_{\ell}. Fix r[M(C)]r\in[M(C)] and r1,,rr_{1},\ldots,r_{\ell} such that r=r1++rr=r_{1}+\ldots+r_{\ell} and rj{0,,μ(Cj)}r_{j}\in\{0,\ldots,\mu(C_{j})\}. For j[]j\in[\ell], let Djrj(Cj)D_{j}\in\mathcal{M}_{r_{j}}(C_{j}) be such that |σj(Dj)|=drj(Cj)|\sigma_{j}(D_{j})|=d_{r_{j}}(C_{j}). Let D=D1××DD=D_{1}\times\ldots\times D_{\ell}. Then Dr(C)D\in\mathcal{M}_{r}(C) has |σ(D)|=dr1(C1)++dr(C)|\sigma(D)|=d_{r_{1}}(C_{1})+\ldots+d_{r_{\ell}}(C_{\ell}), proving that

dr(C)min{j=1drj(Cj):r1++r=r,rj{0,,μ(Cj)}}.d_{r}(C)\leq\min\left\{\sum_{j=1}^{\ell}d_{r_{j}}(C_{j})\,:\,r_{1}+\ldots+r_{\ell}=r,\;r_{j}\in\{0,\ldots,\mu(C_{j})\}\right\}.

To prove the reverse inequality, let D=D1××Dr(C)D=D_{1}\times\ldots\times D_{\ell}\in\mathcal{M}_{r}(C). By Theorem 3.4(1) each of the minimal codewords of CC, say v1,,vrv_{1},\ldots,v_{r}, that minimally generate DD belongs to 0××0×Dj×0××00\times\ldots\times 0\times D_{j}\times 0\times\ldots\times 0 for some j[]j\in[\ell]. Let

rj=|{v1,,vr}(0××0×Dj×0××0)|.r_{j}=|\{v_{1},\ldots,v_{r}\}\cap(0\times\ldots\times 0\times D_{j}\times 0\times\ldots\times 0)|.

Then r=r1++rr=r_{1}+\ldots+r_{\ell} and 0rjμ(Dj)μ(Cj)0\leq r_{j}\leq\mu(D_{j})\leq\mu(C_{j}) for j[]j\in[\ell]. Moreover,

|σ(D)|=j=1|σj(Dj)|j=1drj(Cj),|\sigma(D)|=\sum_{j=1}^{\ell}|\sigma_{j}(D_{j})|\geq\sum_{j=1}^{\ell}d_{r_{j}}(C_{j}),

where the last inequality follows from the fact that Djrj(Cj)D_{j}\in\mathcal{M}_{r_{j}}(C_{j}). By Corollary 3.18

dr(C)\displaystyle d_{r}(C) =min{|σ(D)|:Dr(C)}\displaystyle=\min\{|\sigma(D)|\,:\,D\in\mathcal{M}_{r}(C)\}
min{j=1drj(Cj):r=r1++r,rj{0,,μ(Cj)}}.\displaystyle\geq\min\left\{\sum_{j=1}^{\ell}d_{r_{j}}(C_{j})\,:\,r=r_{1}+\ldots+r_{\ell},\,r_{j}\in\{0,\ldots,\mu(C_{j})\}\right\}.\qed

4 Codes, Supports, and Monomial Ideals

In this section, we prove that the generalized weights of an RR-linear code endowed with a modular support are determined by the graded Betti numbers of a monomial ideal associated to the code. We follow the notation of the previous sections.

Notation 4.1.

In the sequel we work in the multivariate polynomial ring S=K[x1,,xu]S=K[x_{1},\ldots,x_{u}], where KK is an arbitrary field. A monomial of SS is a polynomial of the form x1a1xuaux_{1}^{a_{1}}\cdots x_{u}^{a_{u}}, where (a1,,au)u(a_{1},\ldots,a_{u})\in\mathbb{N}^{u}. In particular, we assume that monomials are monic. A monomial ideal is an ideal which has a system of generators consisting of monomials.

We fix a modular support σ\sigma on RnR^{n}. The support σ\sigma can be used to associate a monomial ideal to a subcode of RnR^{n} as follows.

Definition 4.2.

For any vRn{0}v\in R^{n}\setminus\{0\}, let

mv:=xσ(v):=x1σ(v)1xuσ(v)uS.m_{v}:=x^{\sigma(v)}:=x_{1}^{\sigma(v)_{1}}\cdots x_{u}^{\sigma(v)_{u}}\in S.

For 0CRn0\neq C\subseteq R^{n}, let

IC:=(mv:vC{0})S.I_{C}:=\left(m_{v}\,:\,v\in C\setminus\{0\}\right)\subseteq S.

Notice that not every monomial mICm\in I_{C} corresponds to the support of a codeword vC{0}v\in C\setminus\{0\}. However, every monomial mICm\in I_{C} is of the form m=mvmm=m_{v}\cdot m^{\prime} for some vC{0}v\in C\setminus\{0\} and some monomial mSm^{\prime}\in S.

Proposition 4.3.

Let 0CRn0\neq C\subseteq R^{n} be a code. Then

IC=I0:CJ=(mv:vMin(C))I_{C}=I_{0:_{C}J}=\left(m_{v}\,:\,v\in\textnormal{Min}(C)\right)

and {mv:vMin(C))}\left\{m_{v}\,:\,v\in\textnormal{Min}(C)\right)\} is a minimal system of generators of ICI_{C}.

The next theorem is the main result of this paper. We prove that the graded Betti numbers of the monomial ideal associated to a code determine its generalized weights.

Theorem 4.4.

Let σ\sigma be a modular support and let CRnC\subseteq R^{n} be a code. Assume that either C0:RnJC\subseteq 0:_{R^{n}}J or RR is a PIR. Let ICSI_{C}\subseteq S be the monomial ideal associated to CC and let r[M(C)]r\in[M(C)]. Then M(C)M(C) is the projective dimension of S/ICS/I_{C} and dr(C)d_{r}(C) is the minimum shift (i.e., the minimum degree of a nonzero element) in the rr-th free module in a minimal free resolution of S/ICS/I_{C}. In particular, the graded Betti numbers of S/ICS/I_{C} determine M(C)M(C) and the generalized weights of CC.

Proof.

Let IC=(m1,,mt)I_{C}=(m_{1},\ldots,m_{t}) where m1,,mtm_{1},\ldots,m_{t} are a minimal system of monomial generators of ICI_{C}. By Theorem 3.4(3) v=w\langle v\rangle=\langle w\rangle if and only if σ(v)=σ(w)\sigma(v)=\sigma(w) for v,wMin(C)v,w\in\textnormal{Min}(C). By Theorem 3.5, 0:CJ0:_{C}J has a minimal system of generators consisting of minimal codewords, hence tM(0:CJ)=M(C)t\geq M(0:_{C}J)=M(C) by Proposition 1.10. For each i[t]i\in[t] let viMin(C)v_{i}\in\textnormal{Min}(C) such that mvi=mim_{v_{i}}=m_{i}. For any A[t]A\subseteq[t] let

mA=lcm{mi:iA}=xσ(vi:iA).m_{A}=\operatorname{lcm}\{m_{i}\,:\,i\in A\}=x^{\sigma(\langle v_{i}\,:\,i\in A\rangle)}.

A graded free resolution of S/ICS/I_{C} is given by the Taylor complex [9, Section 7.1]

0𝔽tft𝔽t1ft1f2𝔽1f1SS/IC0,0\longrightarrow\mathbb{F}_{t}\stackrel{{\scriptstyle f_{t}}}{{\longrightarrow}}\mathbb{F}_{t-1}\stackrel{{\scriptstyle f_{t-1}}}{{\longrightarrow}}\ldots\stackrel{{\scriptstyle f_{2}}}{{\longrightarrow}}\mathbb{F}_{1}\stackrel{{\scriptstyle f_{1}}}{{\longrightarrow}}S\longrightarrow S/I_{C}\longrightarrow 0,

where

𝔽r=A[t],|A|=rS(deg(mA))\mathbb{F}_{r}=\bigoplus_{A\subseteq[t],\;|A|=r}S(-\deg(m_{A}))

with basis {eA:A[t],|A|=r}\{e_{A}\,:\,A\subseteq[t],\;|A|=r\} and

fr(eA)=k=1r(1)k+1mAmA{ik}eA{ik} for A={i1,,ir}[t].f_{r}(e_{A})=\sum_{k=1}^{r}(-1)^{k+1}\frac{m_{A}}{m_{A\setminus\{i_{k}\}}}e_{A\setminus\{i_{k}\}}\;\mbox{ for }\;A=\{i_{1},\ldots,i_{r}\}\subseteq[t].

The Taylor resolution is in general not minimal: A cancellation occurs between the modules S(deg(mA))𝔽rS(-\deg(m_{A}))\subseteq\mathbb{F}_{r} and S(deg(mB))𝔽r1S(-\deg(m_{B}))\subseteq\mathbb{F}_{r-1} if and only if B=A{k}B=A\setminus\{k\} for some kAk\in A and mA=mBm_{A}=m_{B}. Notice that mA=mBm_{A}=m_{B} if and only if mklcm{mi:iB}m_{k}\mid\operatorname{lcm}\{m_{i}\,:\,i\in B\}, that is, if and only if σ(vk)iBσ(vi)\sigma(v_{k})\leq\vee_{i\in B}\sigma(v_{i}).

Let

0𝔾pgp𝔾p1gp1g2𝔾1g1SS/IC00\longrightarrow\mathbb{G}_{p}\stackrel{{\scriptstyle g_{p}}}{{\longrightarrow}}\mathbb{G}_{p-1}\stackrel{{\scriptstyle g_{p-1}}}{{\longrightarrow}}\ldots\stackrel{{\scriptstyle g_{2}}}{{\longrightarrow}}\mathbb{G}_{1}\stackrel{{\scriptstyle g_{1}}}{{\longrightarrow}}S\longrightarrow S/I_{C}\longrightarrow 0 (4.1)

be a minimal free resolution obtained from the Taylor resolution after making all the possible cancellations. In particular, pp is the projective dimension of S/ICS/I_{C}.

For A[t]A\subseteq[t] with |A|=r|A|=r, let

CA=vi:iAC.C_{A}=\langle v_{i}\,:\,i\in A\rangle\subseteq C.

Notice that, if vkvi:iA{k}v_{k}\in\langle v_{i}\,:\,i\in A\setminus\{k\}\rangle for some kAk\in A, then S(deg(mA))S(-\deg(m_{A})) cancels with S(deg(mA{k}))S(-\deg(m_{A\setminus\{k\}})) while passing from the Taylor resolution to resolution (4.1). Therefore, the direct summands S(deg(mA))S(-\deg(m_{A})) appearing in (4.1) come from subcodes CAr(C)C_{A}\in\mathcal{M}_{r}(C). Since i(C)=0\mathcal{M}_{i}(C)=0 for i>M(C)i>M(C), then pM(C)p\leq M(C).

For r[M(C)]r\in[M(C)], let brb_{r} be the smallest shift appearing in the rr-th module of a minimal graded free resolution of S/ICS/I_{C}, i.e. brb_{r} is the smallest degree of a nonzero element of 𝔾r\mathbb{G}_{r}. In particular, b1b_{1} is the smallest degree of a minimal generator of ICI_{C}, hence

b1=d1(C).b_{1}=d_{1}(C).

We claim that br=dr(C)b_{r}=d_{r}(C) for all r[M(C)]r\in[M(C)]. Fix a value of rr and choose A[t]A\subseteq[t] with |A|=r|A|=r such that br=deg(mA)b_{r}=\deg(m_{A}). Then σ(CA)=iAσ(vi)=br\sigma(C_{A})=\vee_{i\in A}\sigma(v_{i})=b_{r}. Since CAr(C)C_{A}\in\mathcal{M}_{r}(C), then

dr(C)brd_{r}(C)\leq b_{r} (4.2)

by Theorem 3.14.

To prove the reverse inequality of (4.2), we start by observing that

dr(C)=min{|σ(CA)|:|A|=r,CAr(C)}d_{r}(C)=\min\{|\sigma(C_{A})|\ :\ |A|=r,\ C_{A}\in\mathcal{M}_{r}(C)\}

by Theorem 3.14 and Theorem 3.4(3). Hence dr(C)d_{r}(C) is one of the shifts appearing in the rr-th free module of the Taylor resolution of S/ICS/I_{C}. In order to complete the proof, it suffices to prove the following

Claim A.

𝔾r\mathbb{G}_{r} contains at least one direct summand S(dr(C))S(-d_{r}(C)).

To prove the claim, suppose that we have made all the possible cancellations until the (r1)(r-1)-st step of the resolution. Therefore we have a free resolution of the form

0𝔽tft𝔽r+1hr+1rhr𝔾r1gr1g2𝔾1g1SS/IC0.0\longrightarrow\mathbb{F}_{t}\stackrel{{\scriptstyle f_{t}}}{{\longrightarrow}}\ldots\mathbb{F}_{r+1}\stackrel{{\scriptstyle h_{r+1}}}{{\longrightarrow}}\mathbb{H}_{r}\stackrel{{\scriptstyle h_{r}}}{{\longrightarrow}}\mathbb{G}_{r-1}\stackrel{{\scriptstyle g_{r-1}}}{{\longrightarrow}}\ldots\stackrel{{\scriptstyle g_{2}}}{{\longrightarrow}}\mathbb{G}_{1}\stackrel{{\scriptstyle g_{1}}}{{\longrightarrow}}S\longrightarrow S/I_{C}\longrightarrow 0.

By Theorem 3.12 and Corollary 3.18, r\mathbb{H}_{r} contains a direct summand S(dr(C))S(-d_{r}(C)). Consider now the possible cancellations between 𝔽r+1\mathbb{F}_{r+1} and r\mathbb{H}_{r}. The map

rhrSyzr1(IC)\mathbb{H}_{r}\stackrel{{\scriptstyle h_{r}}}{{\longrightarrow}}\mathrm{Syz}_{r-1}(I_{C}) (4.3)

is surjective and corresponds to a choice of generators of the (r1)(r-1)-st syzygy module Syzr1(IC)\mathrm{Syz}_{r-1}(I_{C}) of ICI_{C}. A cancellation between 𝔽r+1\mathbb{F}_{r+1} and r\mathbb{H}_{r} comes from an element in the kernel of hrh_{r} which has an invertible entry, hence it corresponds to eliminating a non-minimal generator of Syzr1(IC)\mathrm{Syz}_{r-1}(I_{C}). Claim A amounts to showing that, among all direct summands S(dr(C))S(-d_{r}(C)) of r\mathbb{H}_{r}, there is at least one which does not cancel with a direct summand of 𝔽r+1\mathbb{F}_{r+1}. If they all cancel, however, Syzr1(IC)\mathrm{Syz}_{r-1}(I_{C}) has no elements in degree dr(C)d_{r}(C). However this is not possible, since the map in (4.3) is surjective and no component of hrh_{r} is the zero map, hence the image of hrh_{r} contains a nonzero element of degree dr(C)d_{r}(C). In particular, 𝔾p=S(|σ(0:CJ)|)s\mathbb{G}_{p}=S(-|\sigma(0:_{C}J)|)^{s} for some s1s\geq 1 and p=M(C)p=M(C). ∎

Since minimal free resolutions of monomial ideals are easy to compute, Theorem 4.4 gives a way to efficiently compute the generalized weights of a code, if one knows the supports of its minimal codewords. Notice however that the problem of computing the supports of the minimal codewords (or just the minimum distance) of a code is NP-hard, even for the special case of binary linear codes endowed with the Hamming distance, see [16].

We conclude the section with some examples.

Example 4.5.

Let R=6R=\mathbb{Z}_{6} and consider the support ω:62\omega:\mathbb{Z}_{6}\to\mathbb{N}^{2} defined as follows:

ω(0)=(0,0),ω(1)=(2,1),ω(2)=(0,1),ω(3)=(2,0),ω(4)=(0,1),ω(5)=(2,1).\omega(0)=(0,0),\quad\omega(1)=(2,1),\quad\omega(2)=(0,1),\quad\omega(3)=(2,0),\quad\omega(4)=(0,1),\quad\omega(5)=(2,1).

It can be checked that ω\omega is modular. We extend ω\omega componentwise to a function ω:636\omega:\mathbb{Z}_{6}^{3}\to\mathbb{N}^{6} and compose it with the 6\mathbb{Z}_{6}-linear map defined by the invertible matrix

A=(341533245).A=\begin{pmatrix}3&4&1\\ 5&3&3\\ 2&4&5\end{pmatrix}.

In other words, we define σ:636\sigma:\mathbb{Z}_{6}^{3}\to\mathbb{N}^{6} by σ(v)=(ω((Av)1),ω((Av)2),ω((Av)3))\sigma(v)=(\omega((Av)_{1}),\omega((Av)_{2}),\omega((Av)_{3})) for all v63v\in\mathbb{Z}_{6}^{3}. Then σ\sigma is modular by Proposition 2.17(7). Let C=(3,1,2),(2,4,3)63C=\langle(3,1,2),(2,4,3)\rangle\subseteq\mathbb{Z}_{6}^{3}. Then CC has the following six minimal codewords and supports:

(3,3,3)\displaystyle(3,3,3) (0,0,2,0,2,0),\displaystyle(0,0,2,0,2,0),
(0,4,2)\displaystyle(0,4,2) (0,0,0,0,0,1),\displaystyle(0,0,0,0,0,1),
(2,0,4)\displaystyle(2,0,4) (0,1,0,1,0,0),\displaystyle(0,1,0,1,0,0),
(3,3,0)\displaystyle(3,3,0) (2,0,0,0,0,0),\displaystyle(2,0,0,0,0,0),
(0,2,4)\displaystyle(0,2,4) (0,0,0,0,0,1),\displaystyle(0,0,0,0,0,1),
(4,0,2)\displaystyle(4,0,2) (0,1,0,1,0,0).\displaystyle(0,1,0,1,0,0).

In particular, the ideal associated to CC is

IC=(x12,x2x4,x32x52,x6).I_{C}=(x_{1}^{2},x_{2}x_{4},x_{3}^{2}x_{5}^{2},x_{6}).

Notice that the generators of ICI_{C} form a regular sequence, therefore a minimal free resolution of S/ICS/I_{C} is given by the Koszul complex. In other words, there is no cancellation in the Taylor complex

0S(9)S(8)S(7)2S(5)S(6)2S(5)2S(4)S(3)2S(4)S(2)2S(1)SS/IC0.0\longrightarrow S(-9)\longrightarrow\begin{array}[]{c}S(-8)\\ \oplus\\ S(-7)^{2}\\ \oplus\\ S(-5)\end{array}\longrightarrow\begin{array}[]{c}S(-6)^{2}\\ \oplus\\ S(-5)^{2}\\ \oplus\\ S(-4)\\ \oplus\\ S(-3)^{2}\end{array}\longrightarrow\begin{array}[]{c}S(-4)\\ \oplus\\ S(-2)^{2}\\ \oplus\\ S(-1)\end{array}\longrightarrow S\longrightarrow S/I_{C}\longrightarrow 0.

Therefore one has M(C)=4M(C)=4, d1(C)=1d_{1}(C)=1, d2(C)=3d_{2}(C)=3, d3(C)=5d_{3}(C)=5, and d4(C)=9d_{4}(C)=9. Looking at the maps in the minimal free resolutions, one sees that

d1(C)\displaystyle d_{1}(C) =|σ((0,2,4))|=|(0,0,0,0,0,1)|=1,\displaystyle=|\sigma(\langle(0,2,4)\rangle)|=|(0,0,0,0,0,1)|=1,
d2(C)\displaystyle d_{2}(C) =|σ((0,2,4),(3,3,0))|=|σ((0,2,4),(2,0,4))|=|(2,0,0,0,0,1)|=|(0,1,0,1,0,1)|=3,\displaystyle=|\sigma(\langle(0,2,4),(3,3,0)\rangle)|=|\sigma(\langle(0,2,4),(2,0,4)\rangle)|=|(2,0,0,0,0,1)|=|(0,1,0,1,0,1)|=3,
d3(C)\displaystyle d_{3}(C) =|σ((0,2,4),(3,3,0),(2,0,4))|=|(2,1,0,1,0,1)|=5,\displaystyle=|\sigma(\langle(0,2,4),(3,3,0),(2,0,4)\rangle)|=|(2,1,0,1,0,1)|=5,
d4(C)\displaystyle d_{4}(C) =|σ((0,2,4),(3,3,0),(2,0,4),(3,3,3))|=|(2,1,2,1,2,1)|=9.\displaystyle=|\sigma(\langle(0,2,4),(3,3,0),(2,0,4),(3,3,3)\rangle)|=|(2,1,2,1,2,1)|=9.

In the next example, we show that all the cancellations discussed in the proof of Theorem 4.4 can actually occur.

Example 4.6.

Let σ\sigma be the Hamming weight and let C𝔽24C\subseteq{\mathbb{F}}_{2}^{4} be the even weight code, i.e., the code consisting of all vectors of even weight. The minimal codewords of CC are all vectors of weight 22 of 𝔽24{\mathbb{F}}_{2}^{4}, therefore the associated ideal ICI_{C} is the ideal generated by all squarefree monomials of degree 22 in 44 variables. A minimal free resolution of S/ICS/I_{C} is well-known and has the form

0S(4)2S(3)8S(2)6SS/IC0.0\longrightarrow S(-4)^{2}\longrightarrow S(-3)^{8}\longrightarrow S(-2)^{6}\longrightarrow S\longrightarrow S/I_{C}\longrightarrow 0.

The Taylor complex associated to S/ICS/I_{C} is as follows

0S(4)S(4)6S(4)15S(4)16S(3)4S(4)3S(3)12S(2)6SS/IC0.0\longrightarrow S(-4)\longrightarrow S(-4)^{6}\longrightarrow S(-4)^{15}\longrightarrow\begin{array}[]{c}S(-4)^{16}\\ \oplus\\ S(-3)^{4}\end{array}\longrightarrow\begin{array}[]{c}S(-4)^{3}\\ \oplus\\ S(-3)^{12}\end{array}\longrightarrow S(-2)^{6}\longrightarrow S\longrightarrow S/I_{C}\longrightarrow 0.

We use the notation of the proof of Theorem 4.4 when referring to the free modules in the resolutions above. By comparing the two free resolutions, one sees that the modules 𝔽4{\mathbb{F}}_{4}, 𝔽5{\mathbb{F}}_{5} and 𝔽6{\mathbb{F}}_{6} in the Taylor complex completely cancel. Moreover, the free summand S(3)4S(-3)^{4} of 𝔽3{\mathbb{F}}_{3}, corresponding to the minimal shift in its homological degree, cancels, as well as the direct summand S(4)14S(-4)^{14}. Finally, the direct summands S(3)4S(-3)^{4} and S(4)3S(-4)^{3} of 𝔽2{\mathbb{F}}_{2} cancel.

It is easy to find examples of functions that are not supports according to Definition 2.2, and for which the result of Theorem 4.4 does not hold.

Example 4.7.

Let R=4R=\mathbb{Z}_{4}. Let wtL\textnormal{wt}^{\textnormal{L}} denote the Lee weight and let σL\sigma^{\textnormal{L}} be its coordinatewise extension to 43\mathbb{Z}_{4}^{3}. In Example 2.6 we observed that σL\sigma^{\textnormal{L}} is not a support according to Definition 2.2. Consider the code C=(1,1,0),(3,2,1)43C=\langle(1,1,0),(3,2,1)\rangle\subseteq\mathbb{Z}_{4}^{3}. Then

Min(C)={(1,1,0),(3,3,0),(0,1,3),(0,3,1),(1,0,1),(3,0,3)}\textnormal{Min}(C)=\{(1,1,0),(3,3,0),(0,1,3),(0,3,1),(1,0,1),(3,0,3)\}

and IC=(xy,xz,yz)S=K[x,y,z]I_{C}=(xy,xz,yz)\subseteq S=K[x,y,z]. The graded Betti numbers of a minimal free resolution of S/ICS/I_{C} are

0S(3)2S(2)3SS/IC0.0\longrightarrow S(-3)^{2}\longrightarrow S(-2)^{3}\longrightarrow S\longrightarrow S/I_{C}\longrightarrow 0.

In particular, b1=2b_{1}=2 and b2=3b_{2}=3. One can check that d1(C)=4d_{1}(C)=4 and d2(C)=6d_{2}(C)=6.

Notice that Theorem 4.4 does not hold in general for supports that are not modular, even for linear codes over fields.

Example 4.8.

Let C=𝔽22C=\mathbb{F}_{2}^{2} with the support σ\sigma defined in Example 2.21. The associated monomial ideal is I𝔽22=(y)K[x,y]I_{\mathbb{F}_{2}^{2}}=(y)\subseteq K[x,y], whose minimal free resolution is

0S(1)SS/(y)0.0\longrightarrow S(-1)\longrightarrow S\longrightarrow S/(y)\longrightarrow 0.

We have μ(𝔽22)=2\mu(\mathbb{F}_{2}^{2})=2, while the projective dimension of S/(y)S/(y) is 11. In particular, the second free module in a minimal free resolution of S/(y)S/(y) is 0 and it does not determine d2(𝔽22)=|σ(𝔽22)|=2d_{2}(\mathbb{F}_{2}^{2})=|\sigma(\mathbb{F}_{2}^{2})|=2.

5 The Hamming Support

In this section we let R=𝔽qR=\mathbb{F}_{q} and make some observations on the Hamming support σH:𝔽qn{0,1}n\sigma^{\textnormal{H}}:\mathbb{F}_{q}^{n}\to\{0,1\}^{n}, see Example 2.5. For v𝔽qnv\in\mathbb{F}_{q}^{n}, we identify σH(v)\sigma^{\textnormal{H}}(v) with a subset of [n][n].

Several authors study the matroid MCM_{C} associated with the parity-check matrix of a code C𝔽qnC\subseteq\mathbb{F}_{q}^{n}, see e.g. [2, 10]. The code CC, the matroid MCM_{C}, and the corresponding Stanley-Reisner ideal ICI_{C} relate as follows:

  1. (1)

    Let A[n]A\subseteq[n]. Then AMCA\in M_{C} if and only if there is no vC{0}v\in C\setminus\{0\} with σH(v)A\sigma^{\textnormal{H}}(v)\subseteq A.

  2. (2)

    The squarefree monomials in ICI_{C} are exactly those of the form xAx^{A}, where AMCA\notin M_{C}.

  3. (3)

    The circuits of MCM_{C} are the supports of the minimal codewords of CC.

  4. (4)

    The minimal monomial generators of ICI_{C} are in one-to-one correspondence with the circuits of the matroid MCM_{C}.

Therefore,

MC=2[n]{A[n]:Aσ(v) for some vC{0}}=2[n]{A[n]:xAIC}.M_{C}=2^{[n]}\setminus\{A\subseteq[n]\,:\,A\supseteq\sigma(v)\mbox{ for some }v\in C\setminus\{0\}\}=2^{[n]}\setminus\{A\subseteq[n]\,:\,x^{A}\in I_{C}\}. (5.1)

The resolutions of the ideals associated to various classes of codes (most notably, MDS codes, one/two weight codes, Reed-Muller codes) have been studied in [7, 6].

The main result of [10] is Theorem 2, which shows that the generalized Hamming weights of a code C𝔽qnC\subseteq\mathbb{F}_{q}^{n} are determined by the graded Betti numbers of ICI_{C}. Our Theorem 4.4 generalizes [10, Theorem 2] with a different and stand-alone proof strategy that does not rely on any matroid theory results.

A known fact about codes C𝔽qnC\subseteq\mathbb{F}_{q}^{n} endowed with the Hamming metric is that they are generated by their minimal codewords [1, Lemma 2.1.4]. This is a special case of Theorem 3.5 in this paper.

Corollary 5.1.

Every code 0C𝔽qn0\neq C\subseteq\mathbb{F}_{q}^{n} is generated by its minimal codewords.

It is natural to ask whether a code C𝔽qnC\subseteq\mathbb{F}_{q}^{n} is also generated by its codewords of maximal support. The answer to this question is negative in general, as the following simple example shows.

Example 5.2.

Let C=𝔽22C={\mathbb{F}}_{2}^{2}. The only codeword of maximal support is (1,1)(1,1) and (1,1)C\langle(1,1)\rangle\subsetneq C.

Although codes C𝔽qnC\subseteq\mathbb{F}_{q}^{n} are in general not generated by their maximal codewords, for qq sufficiently large most codes are.

Proposition 5.3.

Let 1kn1\leq k\leq n. The proportion of codes in 𝔽qn\mathbb{F}_{q}^{n}, within the kk-dimensional ones, generated by their maximal codewords is at least

(q1)n(qk1)(qn1)qk1(qn1)(qkqk11),\frac{(q-1)^{n}\,(q^{k}-1)-(q^{n}-1)q^{k-1}}{(q^{n}-1)(q^{k}-q^{k-1}-1)},

and the above fraction approaches 1 as q+q\to+\infty.

Proof.

We obtain the result follows from lower bounding the number of kk-dimensional codes with |CAn|>qk1|C\cap A_{n}|>q^{k-1}, where AnA_{n} is the set of vectors in 𝔽qn\mathbb{F}_{q}^{n} of weight nn. To obtain such a lower bound, denote by \mathcal{F} the set of kk-dimensional codes in 𝔽qn\mathbb{F}_{q}^{n} and consider the sum

S:=C|CAn|.S:=\sum_{C\in\mathcal{F}}|C\cap A_{n}|.

We have

||=[nk]q.|\mathcal{F}|=\left[\begin{matrix}n\\ k\end{matrix}\right]_{q}.

We write \mathcal{F} as a disjoint union =′′\mathcal{F}=\mathcal{F}^{\prime}\cup\mathcal{F}^{\prime\prime}, where \mathcal{F}^{\prime} is the set of codes CC\in\mathcal{F} with |CAn|>qk1|C\cap A_{n}|>q^{k-1} and ′′\mathcal{F}^{\prime\prime} the set of codes CC\in\mathcal{F} with |CAn|qk1|C\cap A_{n}|\leq q^{k-1}. We then have

S\displaystyle S =\displaystyle= C|CAn|+C′′|CAn|\displaystyle\sum_{C\in\mathcal{F}^{\prime}}|C\cap A_{n}|+\sum_{C\in\mathcal{F}^{\prime\prime}}|C\cap A_{n}| (5.2)
\displaystyle\leq ||(qk1)+(||||)qk1\displaystyle|\mathcal{F}^{\prime}|\,(q^{k}-1)+(|\mathcal{F}|-|\mathcal{F}^{\prime}|)\,q^{k-1}
=\displaystyle= ||(qkqk11)+||qk1.\displaystyle|\mathcal{F}^{\prime}|\,(q^{k}-q^{k-1}-1)+|\mathcal{F}|\,q^{k-1}.

On the other hand, we can rewrite SS as

S=vAn|{C:Cv}|=|An|[n1k1]q.S=\sum_{v\in A_{n}}|\{C\in\mathcal{F}\,:\,C\ni v\}|=|A_{n}|\,\left[\begin{matrix}n-1\\ k-1\end{matrix}\right]_{q}. (5.3)

Combining (5.2) with (5.3), and using the definition of qq-binomial coefficient, we obtain

||/|||An|qk1(qn1)(qkqk11)qk1qkqk11,|\mathcal{F}^{\prime}|/|\mathcal{F}|\geq|A_{n}|\,\frac{q^{k}-1}{(q^{n}-1)(q^{k}-q^{k-1}-1)}-\frac{q^{k-1}}{q^{k}-q^{k-1}-1},

which is the desired estimate, since |An|=(q1)n|A_{n}|=(q-1)^{n}. ∎



References

  • [1] A. Ashikhmin and A. Barg, Minimal vectors in linear codes, IEEE Transactions on Information Theory 44 (1998), no. 5, 2010–2017.
  • [2] A. Barg, The matroid of supports of a linear code, Applicable Algebra in Engineering, Communication and Computing 8 (1997), no. 2, 165–172.
  • [3] T. Britz, Higher support matroids, Discrete Mathematics 307 (2007), no. 17-18, 2300–2308.
  • [4]  , Code enumerators and Tutte polynomials, IEEE Transactions on Information Theory 56 (2010), no. 9, 4350–4358.
  • [5] N. Gassner, Weight-induced distance functions on /ps\mathbb{Z}/p^{s}\mathbb{Z}-codes, Master Thesis (University of Zurich), 2020, Supervisors: J. Rosenthal and V. Weger.
  • [6] S. R. Ghorpade and R. Ludhani, On the purity of resolutions of Stanley-Reisner rings associated to Reed-Muller codes, arXiv preprint, 2102.00308 (2021).
  • [7] S. R. Ghorpade and P. Singh, Pure resolutions, linear codes, and Betti numbers, Journal of Pure and Applied Algebra 224 (2020), no. 10, 106385.
  • [8] C. Greene, Weight enumeration and the geometry of linear codes, Studies in Applied Mathematics 55 (1976), no. 2, 119–128.
  • [9] J. Herzog and T. Hibi, Monomial Ideals, Springer, 2011.
  • [10] T. Johnsen and H. Verdure, Hamming weights and Betti numbers of Stanley-Reisner rings associated to matroids, Applicable Algebra in Engineering, Communication and Computing 24 (2013), no. 1, 73–93.
  • [11] R. Jurrius and R. Pellikaan, Codes, arrangements and matroids, Algebraic Geometry Modeling in Information Theory, World Scientific, 2013, pp. 219–325.
  • [12] J. MacWilliams and N. Sloane, The Theory of Error-Correcting Codes, North-Holland Mathematical Library, 1977.
  • [13] H. Matsumura, Commutative Ring Theory, Cambridge University Press, 1989.
  • [14] J. G Oxley, Matroid Theory, Oxford University Press, 2006.
  • [15] Alberto Ravagnani, Duality of codes supported on regular lattices, with an application to enumerative combinatorics, Designs, Codes and Cryptography 86 (2018), no. 9, 2035–2063.
  • [16] A. Vardy, The intractability of computing the minimum distance of a code, IEEE Transactions on Information Theory 43 (1997), no. 6, 1757–1766.
  • [17] V. K. Wei, Generalized Hamming weights for linear codes, IEEE Transactions on Information Theory 37 (1991), no. 5, 1412–1418.