Skip to article frontmatterSkip to article content

Minimal Presentations and factorizations

Minimal presentations

In mathematics many algebraic objects are given by means of (free) generators and some relations among them. For numerical semigroups it is sometimes useful to have representation of this form.

Let SS be a numerical semigroup minimally generated by {n1,,ne}\{n_1,\dots,n_e\}. Every element sSs\in S admits an expression of the form s=a1n1++aenes=a_1n_1+\dots+ a_e n_e, and for every (a1,,ae)Ne(a_1,\dots,a_e)\in \mathbb{N}^e, the integer a1n1++aeneSa_1n_1+\dots+a_en_e\in S. Thus, the following (monoid) morphism

φS:NeS, φS(a1,,ae)=a1n1++aene \varphi_S: \mathbb{N}^e \to S,\ \varphi_S(a_1,\dots,a_e)=a_1n_1+\dots+a_en_e

is surjective. There is an isomorphism theorem for monoids, and consequently we have that Ne/kerφS\mathbb{N}^e/\ker \varphi_S is isomorphic to SS as a monoid, where kerϕS={(a,b)Ne:φS(a)=φS(b)}\ker \phi_S=\{ (a,b)\in \mathbb{N}^e : \varphi_S(a)=\varphi_S(b)\}. As a congruence, kerφS\ker \varphi_S admints a system of generators, which is known as a presentation for SS.

A minimal presentation for SS is just a minimal generating system of kerφS\ker \varphi_S.

LoadPackage("num");
true
s:=NumericalSemigroup(2,3);
<Numerical semigroup with 2 generators>
MinimalPresentation(s);
[ [ [ 0, 2 ], [ 3, 0 ] ] ]

Thus 2,3\langle 2,3\rangle can be viewed as the monoid generated by aa and bb, which fulfil the relation 3a=2b3a=2b.

Given sSs\in S, the set φS1(s)\varphi_S^{-1}(s) corresponds to all the expressions that ss migh have in terms of the minimal generators of SS. We will call this set, the set of factorizations of ss, and we will denote it by Z(s)\mathsf{Z}(s).

Factorizations(6,s);
[ [ 3, 0 ], [ 0, 2 ] ]
Factorizations(20,s);
[ [ 10, 0 ], [ 7, 2 ], [ 4, 4 ], [ 1, 6 ] ]

A set ρ generates a congruence σ on Ne\mathbb{N}^e if for any pair (a,b)σ(a,b)\in \sigma there exists a chain a1,,ana_1,\dots,a_n such that

  • a1=aa_1=a, an=ba_n=b,
  • for all ii, there exists ciNec_i\in \mathbb{N}^e and (αi,βi)(\alpha_i,\beta_i) such that (ai,ai+1)=(αi+ci,βi+ci)(a_i,a_{i+1})=(\alpha_i+c_i,\beta_i+c_i) and either (αi,βi)ρ(\alpha_i,\beta_i)\in \rho or (βi,αi)ρ(\beta_i,\alpha_i)\in \rho.

You can see the sequence a1,,ana_1,\dots,a_n as a chain of transformations from aa to bb, and at each step we are applying a relation (or trade) chanbe αi\alpha_i by βi\beta_i (or viceversa).

For instance, in the above example ((10,0),(1,6))kerφS((10,0),(1,6))\in \ker\varphi_S. We start with (10,0)(10,0) and apply the only minimal relation we have on our numerical semigroup: ((10,0),(103,2))σ((10,0),(10-3,2))\in \sigma. If we do this a couple of times more, we obtain the chaing (10,0)(10,0), (7,2)(7,2), (4,4)(4,4), and (1,6)(1,6).

s:=NumericalSemigroup(3,5,7);;
MinimalPresentation(s);
[ [ [ 0, 0, 2 ], [ 3, 1, 0 ] ], [ [ 0, 1, 1 ], [ 4, 0, 0 ] ], [ [ 0, 2, 0 ], [ 1, 0, 1 ] ] ]
Factorizations(20,s);
[ [ 5, 1, 0 ], [ 0, 4, 0 ], [ 1, 2, 1 ], [ 2, 0, 2 ] ]

Assume that we want to check that ((5,1,0),(2,0,2))((5,1,0),(2,0,2)) can be obtained from the minimal relations of SS. Notice that these two factorizations have some common support, in fact (5,1,0)=(3,1,0)+(2,0,0)(5,1,0)=(3,1,0)+(2,0,0) and (2,0,2)=(0,0,2)+(2,0,0)(2,0,2)=(0,0,2)+(2,0,0). Thus, if we find a chain going from (3,1,0)(3,1,0) to (0,0,2)(0,0,2), then by adding (2,0,0)(2,0,0) to all the steps, we obtain a chain connecting our original factorizations. But we already have this chain, since ((0,0,2),(3,1,0))((0,0,2),(3,1,0)) is already in our minimal presentation.

The idea can be extended to factorizations of larger elements in SS. Whenever they have a common “factor”, remove it, and try to connect the new factorizations. This motivates the definition of R\mathcal{R}-classes.

Let XX be a subset of Ne\mathbb{N}^e. We say that x,yXx,y\in X are connected if there exists a sequence x1,,xnXx_1,\dots,x_n\in X (for some nNn\in\mathbb{N}) such that

  • x1=xx_1=x, xn=yx_n=y,
  • for every ii, xix_i and xi+1x_{i+1} have common support.

The connected components of XX under this relation are called RR-classes of XX.

Length(RClassesOfSetOfFactorizations(Factorizations(20,s)));
1

The idea is that whenever we are in the same RR-class, we can connect any two factorizations with a chain where two consequtive elements have common support. For these two elements, we remove the common part, moving now to the factorizations of an element smaller than the original one. If these two new factorizations are in the same RR-class, we repeat the process. This will end when we arrive at factorizations in different RR-classes, and then we only have to connect all the chains that we found translated accordingly.

Let us illustrate this with an example.

Factorizations(21,s);
[ [ 7, 0, 0 ], [ 2, 3, 0 ], [ 3, 1, 1 ], [ 0, 0, 3 ] ]
RClassesOfSetOfFactorizations(Factorizations(21,s));
[ [ [ 0, 0, 3 ], [ 2, 3, 0 ], [ 3, 1, 1 ], [ 7, 0, 0 ] ] ]
JupyterSplashDot(DotEliahouGraph(Factorizations(21,s)));
Loading...

We know that ((7,0,0),(0,0,3))kerφS((7,0,0),(0,0,3))\in \ker\varphi_S, and we want to see how can we obtain this pair from the minimal relations of SS. Both (7,0,0)(7,0,0) and (0,0,3)(0,0,3) are in the same RR-class. We can, instance, connect them with the sequence (7,0,0)(7,0,0), (3,1,1)(3,1,1), (0,0,3)(0,0,3). By transitivity, if we find a chain of trades going from (7,0,0)(7,0,0) to (3,1,1))(3,1,1)) and another from (3,1,1)(3,1,1) to (0,0,3)(0,0,3), by joining them we will find a chain from (7,0,0)(7,0,0) to (0,0,3)(0,0,3).

By removing the common part of (7,0,0)(7,0,0) and (3,1,1)(3,1,1), we obtain (4,0,0)(4,0,0) and (0,1,1)(0,1,1), and the pair ((4,0,0),(0,1,1))((4,0,0),(0,1,1)) is in our minimal presentation. We do the same with (3,1,1)(3,1,1) and (0,0,3)(0,0,3), obtaining (3,1,0)(3,1,0) and (0,0,2)(0,0,2), and ((0,0,2),(3,1,0))((0,0,2),(3,1,0)) is in our minimal presentation. Let σ=kerφS\sigma=\ker\varphi_S. Then (7,0,0)=((4,0,0)+(3,0,0))σ((0,1,1)+(3,0,0))=(3,1,1)σ((3,1,0)+(0,0,1))σ((0,0,2)+(0,0,1))=(0,0,3)(7,0,0)=((4,0,0)+(3,0,0))\sigma ((0,1,1)+(3,0,0))=(3,1,1)\sigma ((3,1,0)+(0,0,1))\sigma ((0,0,2)+(0,0,1))=(0,0,3).

In this construction, elements with more than one RR-class are crucial. These elements are called Betti elements (or degrees) of SS.

bs:=BettiElements(s);
[ 10, 12, 14 ]
List(bs,b-> RClassesOfSetOfFactorizations(Factorizations(b,s)));
[ [ [ [ 0, 2, 0 ] ], [ [ 1, 0, 1 ] ] ], [ [ [ 4, 0, 0 ] ], [ [ 0, 1, 1 ] ] ], [ [ [ 3, 1, 0 ] ], [ [ 0, 0, 2 ] ] ] ]

It is not hard to prove that if sSs\in S is a Betti element, then s=ni+ws=n_i+w with i>2i>2 and wAp(S,n1)w\in \operatorname{Ap}(S,n_1). Thus the number of Betti elements is finite. As a consequence of this, all minimal presentations have the same cardinality, since we only need relations “connecting” different RR-classes of the factorizations of the Betti elements of ss. This also provides a way do determine all minimal relations of a numerical semigroup (up to symmetry).

AllMinimalRelationsOfNumericalSemigroup(s);
[ [ [ 1, 0, 1 ], [ 0, 2, 0 ] ], [ [ 3, 1, 0 ], [ 0, 0, 2 ] ], [ [ 4, 0, 0 ], [ 0, 1, 1 ] ] ]

Minimal presentations of numerical semigroups with embedding dimension two have cardinality one. Thos for embedding dimension three have cardinality two or three. There are numerical semigroups with embedding dimension four with arbitrarily large minimal presentations.

Minimal presentations and binomial ideals

Let SS be a numerical semigroup minimally generated by {n1,,ne}\{n_1,\dots,n_e\}. Let KK be a field and K[x1,,xe]K[x_1,\dots,x_e] be the polynomial ring on the variables x1,,xex_1,\dots, x_e with coefficients in KK.

Let tt be another unknown. We can define the subring K[S]=K[ts:sS]=K[tn1,dots,tne]K[t]K[S]=K[t^s :s \in S]=K[t^{n_1},dots,t^{n_e}]\subseteq K[t], which is known as the semigroup ring of SS. Let ψS:K[x1,,xe]K[t]\psi_S:K[x_1,\dots,x_e]\to K[t] be the unique ring homomorphism determined by ψS(xi)=tni\psi_S(x_i)=t^{n_i}.

For a=(a1,,ae)Nea=(a_1,\dots,a_e)\in \mathbb{N}^e, write xa=x1a1xeaex^a=x_1^{a_1}\cdots x_e^{a_e}. Clearly, ψS(xa)=tϕS(a)\psi_S(x^a)=t^{\phi_S(a)}, and consequently whenever (a,b)kerψS(a,b)\in \ker \psi_S, we have that xaxbkerψSx^a-x^b\in \ker \psi_S. It is not hard to prove that

kerψS=Xaxb:(a,b)kerϕS \ker \psi_S = \langle X^a - x^b : (a,b)\in \ker \phi_S\rangle

and that a set ρ generates kerϕS\ker \phi_S as a congruence if and only if {xaxb:(a,b)ρ}\{ x^a-x^b : (a,b)\in \rho\} generates kerψS\ker\psi_S as an ideal.

s:=NumericalSemigroup(3,5,7);;

In order to compute kerψS\ker \psi_S we can use elimination in the following way.

x:=Indeterminate(Rationals,"x");;
y:=Indeterminate(Rationals,"y");;
z:=Indeterminate(Rationals,"z");;
t:=Indeterminate(Rationals,"t");;

We identify each variable to tt to the power of the corresponding minimal generator.

gens:=[x-t^3,y-t^5,z-t^7];;

Now, we eliminate the variable tt.

gr:=GroebnerBasis(gens,EliminationOrdering([t]));
[ -t^3+x, -t^5+y, -t^7+z, x*t^2-y, x^2*t-z, -x^2+y*t, -x^3+z*t^2, x*y-z*t, -x*z+y^2, -x^4+y*z, -x^3*y+z^2, x^2*y^3-z^3, -x*y^5+z^4, y^7-z^5 ]
gensi:=Filtered(gr, g->DegreeIndeterminate(g,t)=0);
[ -x*z+y^2, -x^4+y*z, -x^3*y+z^2, x^2*y^3-z^3, -x*y^5+z^4, y^7-z^5 ]

Next, we extract the exponents of the binomials obtained.

ed:=3;;
bintopair:=function(pp)
    local m1,m2, d1, d2, p;
    p:=pp/LeadingCoefficientOfPolynomial(pp,MonomialLexOrdering());
    m1:=LeadingMonomialOfPolynomial(p, MonomialLexOrdering());
    m2:=m1-p;
    d1:=List([1..ed], i->DegreeIndeterminate(m1,i));;
    d2:=List([1..ed], i->DegreeIndeterminate(m2,i));;
    return Set([d1,d2]);
end;
function( pp ) ... end
rho:=List(gensi,bintopair);
[ [ [ 0, 2, 0 ], [ 1, 0, 1 ] ], [ [ 0, 1, 1 ], [ 4, 0, 0 ] ], [ [ 0, 0, 2 ], [ 3, 1, 0 ] ], [ [ 0, 0, 3 ], [ 2, 3, 0 ] ], [ [ 0, 0, 4 ], [ 1, 5, 0 ] ], [ [ 0, 0, 5 ], [ 0, 7, 0 ] ] ]

The resulting set does not have to be minimal in general. We can detect those factorizations that correspond to Betti elements.

betticand:=Set(rho, p->p[1]*MinimalGenerators(s));
[ 10, 12, 14, 21, 28, 35 ]
betti:=Filtered(betticand, b->Length(RClassesOfSetOfFactorizations(Factorizations(b,s)))>1);
[ 10, 12, 14 ]
List(betti, b->RClassesOfSetOfFactorizations(Factorizations(b,s)));
[ [ [ [ 0, 2, 0 ] ], [ [ 1, 0, 1 ] ] ], [ [ [ 4, 0, 0 ] ], [ [ 0, 1, 1 ] ] ], [ [ [ 3, 1, 0 ] ], [ [ 0, 0, 2 ] ] ] ]

Length based factorization invariants

Let SS be minimally generated by {n1,,ne}\{n_1,\dots,n_e\}. Recall that the set of factorizations of an element sSs\in S is Z(s)=φS1(s)\mathsf{Z}(s)=\varphi_S^{-1}(s), that is,

Z(s)={(a1,,ae)Ne:a1n1++aene=s}. \mathsf{Z}(s)=\{ (a_1,\dots,a_e)\in \mathbb{N}^e : a_1n_1+\dots+a_en_e =s\}.

Sets of lengths of factorizations

The length of a factorization z=(z1,,ze)z=(z_1,\dots,z_e) is the number of minimal generators involved in it, that is, z=z1++ze|z|=z_1+\dots+z_e. We define the set of lenghts of the factorizations of ss as

L(s)={z:zZ(s)}. \mathsf{L}(s)= \{ |z| : z\in \mathsf{Z}(s)\}.
s:=NumericalSemigroup(3,5,7);;
LengthsOfFactorizationsElementWRTNumericalSemigroup(200,s);
[ 30, 32, 34, 36, 38, 40, 42, 44, 46, 48, 50, 52, 54, 56, 58, 60, 62, 64, 66 ]

Delta sets

One can arrange the lengths of the factorizations of an element ss in the following way L(s)={l1<<lt}\mathsf{L}(s)=\{l_1<\dots<l_t\}. The Delta set of ss is then defined as Δ(s)={l2l1,l3l2,,ltlt1}\Delta(s)=\{l_2-l_1,l_3-l_2,\dots, l_t-l_{t-1}\}.

DeltaSet(200,s);
[ 2 ]

The Delta set of SS is the union of all the delta sets of its elements.

DeltaSet(s);
[ 2 ]

Recall that a minimal presentation of ss was

rho:=MinimalPresentation(s);
[ [ [ 0, 0, 2 ], [ 3, 1, 0 ] ], [ [ 0, 1, 1 ], [ 4, 0, 0 ] ], [ [ 0, 2, 0 ], [ 1, 0, 1 ] ] ]

And that one can “walk” from a factorization to any other factorization of the same element by using this minimal relations.

Set(rho, p->AbsInt(Sum(p[1]-p[2])));
[ 0, 2 ]

So it comes as no surprise that Δ(S)={2}\Delta(S)=\{2\}. In fact, it is not difficult to prove min(Δ(S))=gcd{ab:(a,b)σ}\min(\Delta(S))=\gcd\{ |a-b| : (a,b)\in \sigma\}, with σ any minimal presentation of SS, and that max(Δ(S))=max{Δ(b):bBetti(S)}\max(\Delta(S))=\max\{\Delta(b) : b\in \operatorname{Betti}(S)\}.

The idea behing the minimum is that any possible “jump” will be a linear combination of the “jumps” in the minimal relations. As for the maximum, “jumps” are preserved under translations, and thus the largest “jump” will be achieved between two factorizations with no commom support and in a Betti element.

The structure of Δ(S)\Delta(S) is very well known when SS has embedding dimension two or three, and even when it is generated by an arithmetice sequence.

The elasticity

The elasticity of the factorizations of an element ss in a numerical semigroup SS is ratio between the largest length and smallest lenght of its factorizations, that is,

ρ(s)=max(L(s))min(L(s)). \rho(s)=\frac{\max(\mathsf{L}(s))}{\min(\mathsf{L}(s))}.

The elasticity of the semigroup SS is defined as

ρ(S)=sup{ρ(s):sS}. \rho(S)=\sup\{ \rho(s) : s\in S\}.

For numerical semigroups, this supremum becomes a maximum, and the elasticity is attained at the element n1nen_1 n_e (the product of the smallest generator times the largest generator).

Elasticity(100,s);
2
Elasticity(s);
7/3
Elasticity(3*7,s);
7/3

Distance based factorization invariants

Given x=(x1,,xe)x=(x_1,\dots,x_e) and y=(y1,,ye)y=(y_1,\dots,y_e) in Ne\mathbb{N}^e, their “common part” is

xy=(min{x1,y1},,min{xe,ye}), x\wedge y = (\min\{x_1,y_1\},\dots,\min\{x_e,y_e\}),

and the distance between xx and yy as

d(x,y)=max{x(xy),y(xy)}. \operatorname{d}(x,y)=\max\{|x-(x\wedge y)|, |y-(x\wedge y)|\}.
s:=NumericalSemigroup(10,13,19,21);;

We have already used DotEliahouGraph, the labels of the edges are the distances between de factorizations they connect.

JupyterSplashDot(DotEliahouGraph(Factorizations(100,s)));
Loading...

The catenary degree

In addition, DotFactorizationGraph draws a minimum spanning tree.

JupyterSplashDot(DotFactorizationGraph(Factorizations(100,s)));
Loading...

In particular, this means that we can go from any factorization of 100 in SS to any other factorization of the same element by using a path such that two consecutive nodes are at a distance of at most four. This is precisely the idea behind the concept of catenary degree.

Let zz and zz' be two factorizations of ss. An NN-chain joining zz and zz' is a sequence z1,,ztz_1,\dots,z_t of factorizations of ss such that d(zi,zi+1)N\operatorname{d}(z_i,z_{i+1})\le N. The catenary degree of ss, denoted c(s)\operatorname{c}(s), in the minimum NN such that for any two factorizations of ss there exists an NN-chain connecting them.

The catenary degree of SS is

c(S)=sup{c(s):sS}. \operatorname{c}(S)=\sup\{ \operatorname{c}(s) : s\in S\}.

This supremum is a maximum.

CatenaryDegree(100,s);
4
CatenaryDegree(s);
4

Recall that by using the minimal relations of SS we can find a path joining any two different factorizations of an element. Thus, it is not hard to prove that the maximum of the catenary degree of SS is attained at one of its Betti elements.

Set(BettiElements(s), b->CatenaryDegree(b,s));
[ 3, 4 ]

The tame degree

The catenary degree measures the minimum distance needed to find paths connecting any two factorizations of an element in the semigroup in such a way that every step in the path is withing that minimum distance.

The tame degree intends to measure a radius rr in which for any factorization zz of sSs\in S, with sniSs-n_i\in S, you will find another factorization zz' such that d(z,z)r\operatorname{d}(z,z')\le r and zi0z_i'\neq 0.

Let sSs\in S such that sniSs-n_i\in S for some i{1,,e}i\in\{1,\dots,e\}. Define Zi(s)={z:zZ(s),zi0}\mathsf{Z}_i(s)=\{z : z\in \mathsf{Z}(s), z_i\neq 0\}, which is nonempty as sniSs-n_i\in S. Set

ti(s)=sup{d(z,Zi(s)):zZ(n)}, \operatorname{t}_i(s)=\sup\{\operatorname{d}(z,\mathsf{Z}_i(s)) : z \in \mathsf{Z}(n)\},

and

t(s)=max{ti(s):sniS,i{1,,e}}. \operatorname{t}(s)=\max\{\operatorname{t}_i(s) : s-n_i\in S, i\in\{1,\dots,e\}\}.

The tame degree of SS is defined as

t(S)=sup{t(s):sS}. \operatorname{t}(S)=\sup\{ \operatorname{t}(s) : s\in S\}.
s:=NumericalSemigroup( 24, 59, 114);
<Numerical semigroup with 3 generators>
TameDegree(s);
29

Usually the tame degre is not attained at the Betti elements.

Set(BettiElements(s),b->TameDegree(b,s));
[ 11, 19 ]

It can be shown that the tame degree of SS is attained at an element sSs\in S such that ss has a factorization in Minimals(Z(ni+S))\operatorname{Minimals}_\le (\mathsf{Z}(n_i+S)) for some i{1,,e}i\in\{1,\dots,e\}. Elements having this property are of the form ni+wn_i+w with wAp(S,nj)w\in \operatorname{Ap}(S,n_j), more specifically, there exists i,j{1,,e}i,j\in \{1,\dots,e\} such that sni,snjSs-n_i,s-n_j\in S and s(ni+nj)∉Ss-(n_i+n_j)\not\in S.

For sSs\in S, the Rosales graph GsG_s is defined as follows. The vertices of GsG_s are the minimal generators nin_i such that sniSs-n_i\in S, and ninjn_in_j is an edge whenever s(ni+nj)Ss-(n_i+n_j)\in S.

JupyterSplashDot(DotRosalesGraph(400,s));
Loading...
BettiElements(s);
[ 354, 456 ]
JupyterSplashDot(DotRosalesGraph(456,s));
Loading...
JupyterSplashDot(DotRosalesGraph(354,s));
Loading...

The number of connected components of GsG_s coincides with the set of RR-classes of the set of factorizations of ss. Thus the catenary degree is attained in an ss with GsG_s not connected, and the tame degree in an ss with GsG_s not complete.

Primality

Recall that SS induces an order over the integers aSba\le_S b if baSb-a\in S. If aa and bb are in SS, then we say that aa divides bb if baSb-a\in S. In this way, minimal generators (irreducibles, atoms, primitive elements) are those not having proper divisors. A natural question arises: are there “prime” elements in a numerical semigroup. The ones to be candidates to be prime are the minimal generators of the semigroup.

Let SS be minimally generated by {n1,,ne}\{n_1,\dots,n_e\}. Consider the set Minimals(Z(ni+S))\operatorname{Minimals}_\le (\mathsf{Z}(n_i+S)). Notice that S(ni+S)=Ap(S,ni)S\setminus(n_i+S)=\operatorname{Ap}(S,n_i), and so the set Z(ni+S)=NeZ(Ap(S,ni))\mathsf{Z}(n_i+S)=\mathbb{N^e}\setminus \mathsf{Z}(\operatorname{Ap}(S,n_i)). Let zMinimals(Z(ni+S))z \in \operatorname{Minimals}_\le (\mathsf{Z}(n_i+S)), zeiz\neq e_i. Then nin_i divides φS(z)\varphi_S(z) and cannot divide any of its “factors”. Hence, nin_i cannot be prime.

Let sSs\in S. The ω-primality of SS, ω(S,s)\omega(S,s), is defined as the least integer NN such that whenever ss divides a1++ana_1+\dots+a_n for some a1,,anSa_1,\dots,a_n\in S, then ss divides ai1++aiNa_{i_1}+\dots+a_{i_N} for some {i1,,iN}{1,,n}\{i_1,\dots,i_N\}\subseteq \{1,\dots,n\}. By an argument similar to the one given above,

ω(S,s)=max{z:zMinimals(Z(s+S))}. \omega(S,s)= \max \{ |z| : z\in \operatorname{Minimals}_\le (\mathsf{Z}(s+S))\}.
s:=NumericalSemigroup(3,5,7);
<Numerical semigroup with 3 generators>
OmegaPrimality(10,s);
5

Observe that if zMinimals(Z(s+S))z\in \operatorname{Minimals}_\le(\mathsf{Z}(s+S)), then φ(z)=s+t\varphi(z)=s+t for some tSt\in S. Let ii be such zi0z_i\neq 0. Then zzi∉Z(s+S)z-z_i\not\in \mathsf{Z}(s+S), and thus s+tni∉s+Ss+t-n_i\not\in s+S, which means that tAp(S,ni)t\in \operatorname{Ap}(S,n_i). This bounds the search for cmputing Minimals(Z(s+S))\operatorname{Minimals}_\le(\mathsf{Z}(s+S)).

There is an alternative way to compute the set Minimals(Z(s+S))\operatorname{Minimals}_\le(\mathsf{Z}(s+S)). Observe that xZ(s+S)x\in \mathsf{Z}(s+S) if an only if n1x1++xene=s+n1t1++neten_1x_1+\dots+x_en_e=s+n_1t_1+\dots+n_et_e for some t1,,teNt_1,\dots,t_e\in \mathbb{N}. So we can solve the problem by looking at the nonnegative integer solutions (in N2e\mathbb{N}^{2e} to the equation

n1x1++xenen1t1nete=sn_1x_1+\dots+x_en_e-n_1t_1-\dots-n_et_e=s

and then project onto the first ee coordinates.

The omega primality of SS is defined as the maximum of the omega primalities of its minimal generators.

OmegaPrimality(s);
4

It can be shown that

2+max(δ(S))c(S)ω(S)t(S). 2+\max(\delta(S))\le \operatorname{c}(S)\le \omega(S)\le \operatorname{t}(S).

Divisors

Let ss be an element in a numerical semigroup SS. The set of divisors of ss is

D(s)={nS:snS}. \operatorname{D}(s)=\{ n\in S : s-n\in S\}.

The Feng-Rao distance of ss is defined as

δFR(s)=min{#D(s):ss,sS}. \delta_{FR}(s)=\min\{ \# \operatorname{D}(s') : s\le s', s'\in S\}.

Let cc be the conductor of SS and let gg be its genus. It can be shown that for s2c1s\ge 2c-1,

δFR(s)=s+12g. \delta_{FR}(s)=s+1-2g.
ndiv:={x,s}->Length(DivisorsOfElementInNumericalSemigroup(x,s));;
s:=NumericalSemigroup(3,5,7);;
List(s{[1..20]}, x->ndiv(x,s));
[ 1, 2, 2, 3, 2, 4, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17 ]
FengRaoDistance(s,1,10);
5

The generalized Feng-Rao distance is defined as

δFRr=min{#D(s1,,sr):ss1sr,s1,,srS}, \delta_{FR}^r=\min\{\#\operatorname{D}(s_1,\dots,s_r) : s\le s_1\le \dots \le s_r, s_1,\dots,s_r\in S\},

where D(s1,,sr)=i=1rD(si)\operatorname{D}(s_1,\dots,s_r)=\bigcup_{i=1}^r \operatorname{D}(s_i). It can be shown that for s2c1s\ge 2c-1,

δFRr=s+12g+E(S,r). \delta_{FR}^r= s+1 -2g + E(S,r).

The constant E(S,r)E(S,r) is known as the rrth Feng-Rao number. For r=2r=2,

E(S,2)=min{#Ap(S,n):r{1,,m(S)}}. E(S,2)=\min\{\#\operatorname{Ap}(S,n) : r\in \{1,\dots,\operatorname{m}(S)\}\}.

The definition of the Apéry set for elements not in SS is the same as the one given above.