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

The Automorphism groups of zero-dimensional monomial algebras

Roberto Díaz Departamento de Matemáticas, Facultad de Ciencias, Universidad de Chile, Las Palmeras 3425, Ñuñoa, Santiago, Chile. robertodiaz@uchile.cl Alvaro Liendo Instituto de Matemática y Física, Universidad de Talca, Casilla 721, Talca, Chile. aliendo@utalca.cl Gonzalo Manzano-Flores Departamento de Matemáticas, Facultad de Ciencias, Universidad de Chile, Las Palmeras 3425, Ñuñoa, Santiago, Chile. gonzalo.manzano@usach.cl  and  Andriy Regeta Institut für Mathematik, Friedrich-Schiller-Universität Jena, Jena 07737, Germany andriyregeta@gmail.com
Abstract.

A monomial algebra BB is defined as a quotient of a polynomial ring by a monomial ideal, which is an ideal generated by a finite set of monomials. In this paper, we determine the automorphism group of a monomial algebra BB, under the assumption that BB is a finite-dimensional vector space over a field 𝐤\mathbf{k} of characteristic zero. We achieve this by providing an explicit classification of the homogeneous locally nilpotent derivations of BB. The main body of the paper addresses the more general case of semigroup algebras, with the polynomial ring being a particular case.

2020 Mathematics Subject Classification: 13F55, 13N15, 14L30, 14M25.
    Key words: monomial ideal, locally nilpotent derivation, automorphism group.
     The first author is supported by the Fondecyt postdoc project 3230406. The second author was partially supported by Fondecyt project 1240101. The third author is supported by the Fondecyt postdoc project 3240191. The fourth author is supported by DFG, project number 509752046.

Introduction

An affine semigroup SS is a finitely generated semigroup with an identity element that can be embedded into n\mathbb{Z}^{n}. Let 𝐤\mathbf{k} be a field of characteristic zero and SS an affine semigroup. The semigroup algebra 𝐤[S]\mathbf{k}[S] is defined as 𝐤[S]=mS𝐤𝐱m\mathbf{k}[S]=\bigoplus_{m\in S}\mathbf{k}\cdot\mathbf{x}^{m}, where 𝐱m\mathbf{x}^{m} are new variables satisfying the multiplication rules 𝐱m𝐱m=𝐱m+m\mathbf{x}^{m}\cdot\mathbf{x}^{m^{\prime}}=\mathbf{x}^{m+m^{\prime}} and 𝐱0=1\mathbf{x}^{0}=1. A common example of a semigroup algebra, which motivates our notation, is the polynomial ring 𝐤[𝐱]=𝐤[x1,,xn]\mathbf{k}[\mathbf{x}]=\mathbf{k}[x_{1},\ldots,x_{n}] obtained with S=0nS=\mathbb{Z}_{\geq 0}^{n}, where 𝐱m\mathbf{x}^{m} corresponds precisely to monomials under the usual multi-index notation. By analogy, we refer to all symbols 𝐱m\mathbf{x}^{m} as monomials of 𝐤[S]\mathbf{k}[S]. An ideal I𝐤[S]I\subseteq\mathbf{k}[S] is called monomial if it has a generating set composed of monomials. A monomial algebra 𝐤[S]/I\mathbf{k}[S]/I is the quotient of 𝐤[S]\mathbf{k}[S] by a monomial ideal II. A monomial algebra 𝐤[S]/I\mathbf{k}[S]/I is zero-dimensional if and only if its dimension as a 𝐤\mathbf{k}-vector space is finite.

The combinatorial nature of monomial algebras allows them to be identified with objects from other categories, making these algebras suitable for developing examples and verifying general theories. For instance, Stanley-Reisner rings, which are monomial algebras characterized by having all their exponents being 0 or 1, are in bijective correspondence with abstract simplicial complexes [MS05, Theorem 1.7]. Stanley proved the upper bound conjecture for spheres using the Cohen-Macaulay structure of these rings [Sta75], which was established by Reisner [Rei76]. In another example, a special class of monomial algebras known as discrete Hodge algebras was studied in [DCEP82], enabling the examination of homogeneous coordinate rings of Grassmann varieties and their Schubert subvarieties.

Our main goal in this paper is to compute the automorphism groups of zero-dimensional monomial algebras.

Our approach to computing the automorphism group of a zero-dimensional monomial algebra is inspired by Demazure’s foundational paper [Dem70], where he computes the automorphism groups of proper smooth toric varieties. A normal variety is called toric if it admits an algebraic torus action with an open orbit, and affine toric varieties are precisely given by Spec𝐤[S]\operatorname{Spec}\mathbf{k}[S], where SS is an affine semigroup [Oda83, Ful93, CLS11]. Although the geometric nature of smooth proper toric varieties is significantly different from that of zero-dimensional monomial algebras, both objects have similar combinatorial descriptions. This similarity allows us to apply some of the techniques from [Dem70] to our context. We refer the reader to the paper [LLA22] by the second named author for a concise modern account of Demazure’s results.

For a 𝐤\mathbf{k}-algebra BB, we denote by Der(B)\operatorname{Der}(B) the 𝐤\mathbf{k}-vector space of 𝐤\mathbf{k}-derivations :BB\partial\colon B\to B. Additionally, if IBI\subset B is an ideal, we denote by DerI(B)\operatorname{Der}_{I}(B) the subspace of Der(B)\operatorname{Der}(B) consisting of 𝐤\mathbf{k}-derivations such that (I)I\partial(I)\subset I. All our derivations are 𝐤\mathbf{k}-derivations, so we will omit 𝐤\mathbf{k} from the notation. A derivation Der(B)\partial\in\operatorname{Der}(B) is called locally nilpotent if for every fBf\in B there exists 0\ell\in\mathbb{Z}_{\geq 0} such that (f)=0\partial^{\ell}(f)=0, where \partial^{\ell} denotes the \ell-th composition of \partial.

As a first step towards our goal of computing the automorphism groups of monomial algebras, we describe the set DerI(𝐤[S])\operatorname{Der}_{I}(\mathbf{k}[S]), where SS is an affine semigroup and II is a monomial ideal (see Proposition 1.3). This description has already been provided for the special case where the semigroup algebra 𝐤[S]\mathbf{k}[S] is a polynomial ring in [BS95, Tad09]. Consider an embedding SMS\hookrightarrow M where MnM\simeq\mathbb{Z}^{n} such that the group generated by SS in MM is MM itself. The algebra 𝐤[S]\mathbf{k}[S] admits an MM-grading, and since II is MM-graded, 𝐤[S]/I\mathbf{k}[S]/I inherits this MM-grading. We make extensive use of these MM-gradings throughout the paper, and in particular, we derive Proposition 1.3 straightforwardly from [KLL15] by the second named author.

Let SS be a monomial algebra and II be a monomial ideal. There is a natural map π:DerI(𝐤[S])Der(𝐤[S]/I)\pi\colon\operatorname{Der}_{I}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I) induced by the quotient. In Theorem 2.2, we show that π\pi is surjective for every monomial ideal II if and only if 𝐤[S]\mathbf{k}[S] is a polynomial ring. A derivation ¯:𝐤[S]/I𝐤[S]/I\overline{\partial}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I is said to be liftable if there exists a derivation :𝐤[S]𝐤[S]\partial\colon\mathbf{k}[S]\to\mathbf{k}[S] such that π()=¯\pi(\partial)=\overline{\partial}. In Theorem 3.7, we provide a classification of the liftable locally nilpotent derivations on the monomial algebra 𝐤[S]/I\mathbf{k}[S]/I that are homogeneous with respect to the MM-grading.

The interest in locally nilpotent derivations of an algebra BB stems from their one-to-one correspondence with additive group actions on BB; see, for instance, [Dai03, Section 4]. Specifically, given a locally nilpotent derivation :BB\partial\colon B\to B, the associated 𝔾a\mathbb{G}_{\mathrm{a}}-action on BB is obtained via the exponential map:

exp(s):𝔾a×BBdefined by(s,f)i=0sii(f)i!.\displaystyle\exp(s\partial)\colon\mathbb{G}_{\mathrm{a}}\times B\to B\quad\text{defined by}\quad(s,f)\mapsto\sum_{i=0}^{\infty}\frac{s^{i}\partial^{i}(f)}{i!}\,. (1)

To describe our main result, we let S=0nS=\mathbb{Z}_{\geq 0}^{n} so that B=𝐤[S]B=\mathbf{k}[S] is a polynomial ring and we let II be a monomial ideal. Without loss of generality, we may assume that II is full, i.e., no standard basis vector in n\mathbb{Z}^{n} is contained in II (see Definition 4.2). Our main result, presented in Theorem 4.17, implies the following.

Theorem.

Let S=0nS=\mathbb{Z}^{n}_{\geq 0}, making B=𝐤[S]B=\mathbf{k}[S] a polynomial ring, and let IBI\subset B be a full monomial ideal such that B/IB/I is a zero-dimensional monomial algebra. If G=Aut𝐤(B/I)G=\operatorname{Aut}_{\mathbf{k}}(B/I), then the following statements hold:

  1. (i)(i)

    The group GG is linear algebraic.

  2. (ii)(ii)

    The algebraic torus T=Spec𝐤[n]T=\operatorname{Spec}\mathbf{k}[\mathbb{Z}^{n}] is a maximal torus of GG.

  3. (iii)(iii)

    The connected component G0G^{0} of GG is generated by the maximal torus and the images in G0G^{0} of all the 𝔾a\mathbb{G}_{\mathrm{a}}-actions corresponding to homogeneous locally nilpotent derivations :B/IB/I\partial\colon B/I\to B/I.

  4. (iv)(iv)

    The finite group G/G0G/G^{0} is generated by the images of automorphisms B/IB/IB/I\to B/I induced by semigroup automorphisms SSS\to S that map II to itself.

Let now BB be an affine domain, and assume that every automorphism of the connected component of the identity Aut𝐤0(B)\operatorname{Aut}_{\mathbf{k}}^{0}(B) of Aut𝐤(B)\operatorname{Aut}_{\mathbf{k}}(B) is algebraic (see [PR24, page 3] for the definition of an algebraic element). Then, Aut𝐤0(B)\operatorname{Aut}_{\mathbf{k}}^{0}(B) is a solvable group with a derived length of at most two (see [PR24, Theorem 1.1]; see also [PR23, Theorem 1.3]). In contrast, the automorphism group of a zero-dimensional monomial algebra BB does not necessarily have a solvable connected component (see Example 4.20).

The article is organized as follows: In Section 1, we present preliminaries on semigroup algebras, monomial ideals, and homogeneous derivations. Section 2 explores the relationship between derivations on a semigroup algebra and its quotient monomial algebras. In Section 3, we classify liftable locally nilpotent derivations of monomial algebras. Finally, in Section 4, we describe the automorphism group of zero-dimensional monomial algebras that are quotients of the polynomial ring.

Acknowledgments

This collaboration began at the AGREGA2 conference, which took place at Universidad Técnica Federico Santa María in March 2023. We extend our gratitude to the institution for its support and hospitality.

1. Preliminaries

In this section, we gather some preliminary results that are essential for the subsequent sections.

1.1. Semigroup algebras, monomial ideals and homogeneous derivations

Throughout this article, we fix a free abelian group MM of rank nn, i.e., MnM\simeq\mathbb{Z}^{n}. We also consider NN as the dual group, defined by N=Hom(M,)N=\operatorname{Hom}(M,\mathbb{Z}). There exists a natural duality pairing ,:M×N\langle\ ,\ \rangle\colon M\times N\to\mathbb{Z} given by m,p=p(m)\langle m,p\rangle=p(m). Let LL be any field. We define the LL-vector spaces ML=MLM_{L}=M\otimes_{\mathbb{Z}}L and NL=NLN_{L}=N\otimes_{\mathbb{Z}}L. The aforementioned duality pairing extends naturally to a pairing ,:ML×NLL\langle\ ,\ \rangle\colon M_{L}\times N_{L}\to L.

An affine semigroup SS is a finitely generated semigroup with an identity element that can be embedded into MM. The cone ωS\omega_{S} of SS is the cone spanned by SS inside MM_{\mathbb{R}}. We say that an affine semigroup SS is saturated if ωSM=S\omega_{S}\cap M=S, and pointed if the only vector space contained in ωS\omega_{S} is {0}\{0\}. Furthermore, SS is said to be minimally embedded if the smallest subgroup of MM containing SS is MM. Unless otherwise stated, all semigroups in the sequel are assumed to be affine, saturated, pointed, and minimally embedded.

Let SS be a semigroup. We say that a subsemigroup SSS^{\prime}\subset S is a face of SS if for all m,mSm,m^{\prime}\in S with m+mSm+m^{\prime}\in S^{\prime}, we have m,mSm,m^{\prime}\in S^{\prime}. Note that SSS^{\prime}\subset S is a face if and only if S=SτS^{\prime}=S\cap\tau, where τ\tau is a polyhedral face of ωS\omega_{S}. A face of SS isomorphic to 0\mathbb{Z}_{\geq 0} as a semigroup is called a ray, and a maximal proper face of SS is called a facet. The rays of SS are given by SτS\cap\tau with τ\tau a one-dimensional polyhedral face of ωS\omega_{S}, and the facets of SS are given by SτS\cap\tau with τ\tau a (n1)(n-1)-dimensional polyhedral face of ωS\omega_{S}. A ray of SS is uniquely determined by its unique generator as a semigroup; thus, we denote by S(1)SS(1)\subset S the set of all these generators of all the rays. The rank of a face FSF\subset S, denoted by rankF\operatorname{rank}F, is the rank of the subgroup of MM generated by FF. A semigroup is called simplicial if S(1)S(1) is an \mathbb{R}-basis of MM_{\mathbb{R}}.

Let SS be a semigroup. We say that a subsemigroup SSS^{\prime}\subset S is a face of SS if for all m,mSm,m^{\prime}\in S with m+mSm+m^{\prime}\in S^{\prime}, we have m,mSm,m^{\prime}\in S^{\prime}. Note that SSS^{\prime}\subset S is a face if and only if S=SτS^{\prime}=S\cap\tau, where τ\tau is a polyhedral face of ωS\omega_{S}. A face of SS isomorphic to 0\mathbb{Z}_{\geq 0} as a semigroup is called a ray, and a maximal proper face of SS is called a facet. The rays of SS are given by SτS\cap\tau where τ\tau is a one-dimensional polyhedral face of ωS\omega_{S}, and the facets of SS are given by SτS\cap\tau where τ\tau is a (n1)(n-1)-dimensional polyhedral face of ωS\omega_{S}. A ray of SS is uniquely determined by its unique generator as a semigroup. We denote by S(1)SS(1)\subset S the set of all these generators of all the rays. The rank of a face FSF\subset S, denoted by rankF\operatorname{rank}F, is the rank of the subgroup of MM generated by FF. A semigroup is called simplicial if S(1)S(1) is an \mathbb{R}-basis of MM_{\mathbb{R}}.

The semigroup algebra 𝐤[S]\mathbf{k}[S] of SS is defined as:

𝐤[S]=mS𝐤𝐱mwhere𝐱m𝐱m=𝐱m+mand𝐱0=1.\mathbf{k}[S]=\bigoplus_{m\in S}\mathbf{k}\cdot\mathbf{x}^{m}\quad\mbox{where}\quad\mathbf{x}^{m}\cdot\mathbf{x}^{m^{\prime}}=\mathbf{x}^{m+m^{\prime}}\quad\mbox{and}\quad\mathbf{x}^{0}=1\,.

The symbols 𝐱m\mathbf{x}^{m}, with mMm\in M are called monomials. Note that since SS is affine and saturated, 𝐤[S]\mathbf{k}[S] is a normal affine domain.

An ideal I𝐤[S]I\subset\mathbf{k}[S] is called monomial if there exists l>0l\in\mathbb{Z}_{>0} and 𝐚1,,𝐚lS\mathbf{a}_{1},\ldots,\mathbf{a}_{l}\in S such that I=(𝐱𝐚1,,𝐱𝐚l)I=(\mathbf{x}^{\mathbf{a}_{1}},\dots,\mathbf{x}^{\mathbf{a}_{l}}). The support of a monomial ideal I𝐤[S]I\subset\mathbf{k}[S] is denoted by supp(I),\operatorname{supp}(I), and is given by

supp(I)={mS𝐱mI}=k=1l(𝐚k+S).\operatorname{supp}(I)=\{m\in S\mid\mathbf{x}^{m}\in I\}=\bigcup_{k=1}^{l}(\mathbf{a}_{k}+S)\,.

Let I𝐤[S]I\subset\mathbf{k}[S] be a monomial ideal. We say that II has cofinite support if Ssupp(I)S\setminus\operatorname{supp}(I) is finite.

Example 1.1 (The polynomial ring).

Let M=nM=\mathbb{Z}^{n} and let E={e1,,en}E=\{e_{1},\ldots,e_{n}\} be the canonical basis. Let S=0nS=\mathbb{Z}_{\geq 0}^{n}, there exists a natural isomorphism from 𝐤[S]\mathbf{k}[S] to the polynomial ring 𝐤[x1,,xn]\mathbf{k}[x_{1},\ldots,x_{n}], given by 𝐱eixi\mathbf{x}^{e_{i}}\mapsto x_{i}, for 1in1\leq i\leq n. Under this isomorphism, for every m=(m1,,mn)Sm=(m_{1},\ldots,m_{n})\in S, the monomial 𝐱m\mathbf{x}^{m} corresponds to x1m1xnmnx_{1}^{m_{1}}\cdots x_{n}^{m_{n}}.

Let M=nM=\mathbb{Z}^{n} and let E={e1,,en}E=\{e_{1},\ldots,e_{n}\} be the canonical basis. For S=0nS=\mathbb{Z}_{\geq 0}^{n}, there exists a natural isomorphism from 𝕂[S]\mathbb{K}[S] to the polynomial ring 𝕂[x1,,xn]\mathbb{K}[x_{1},\ldots,x_{n}], given by 𝐱eixi\mathbf{x}^{e_{i}}\mapsto x_{i} for 1in1\leq i\leq n. Under this isomorphism, for every m=(m1,,mn)Sm=(m_{1},\ldots,m_{n})\in S, the monomial 𝐱m\mathbf{x}^{m} corresponds to x1m1xnmnx_{1}^{m_{1}}\cdots x_{n}^{m_{n}}.

Now, fixing n=2n=2 we have that 𝐤[S]=𝐤[x,y]\mathbf{k}[S]=\mathbf{k}[x,y]. Let I1=(x2y5,x3y2,x5)𝐤[S]I_{1}=(x^{2}y^{5},x^{3}y^{2},x^{5})\subset\mathbf{k}[S] and I2=(y5,x3y2,x6)𝐤[S]I_{2}=(y^{5},x^{3}y^{2},x^{6})\subset\mathbf{k}[S]. Note that I2I_{2} is a monomial ideal with cofinite support while I1I_{1} is not. The ideal I1I_{1} is represented in Figure 2 and I2I_{2} is represented in Figure 2. The solid dots \bullet represent elements in supp(Ii),\operatorname{supp}(I_{i}), while the hollow dots \circ represent elements in mSsupp(Ii)m\in S\setminus\operatorname{supp}(I_{i}), for 1i21\leq i\leq 2.

Figure 1. I1=(x2y5,x3y2,x5)I_{1}=(x^{2}y^{5},x^{3}y^{2},x^{5})
Figure 2. I2=(y5,x3y2,x6)I_{2}=(y^{5},x^{3}y^{2},x^{6})

The algebra 𝐤[S]\mathbf{k}[S] is naturally MM-graded, and under this grading every monomial ideal II is a graded ideal. Thus, the quotient 𝐤[S]/I\mathbf{k}[S]/I inherits an MM-grading given by deg𝐱¯m=m\deg{\overline{\mathbf{x}}}^{m}=m if msupp(I)m\notin\operatorname{supp}(I), where 𝐱¯m{\overline{\mathbf{x}}}^{m} denotes the image of 𝐱m\mathbf{x}^{m} within 𝐤[S]/I\mathbf{k}[S]/I. Recall that if msupp(I)m\in\operatorname{supp}(I), then 𝐱¯m=0{\overline{\mathbf{x}}}^{m}=0. In the sequel, we will consider 𝐤[S]/I\mathbf{k}[S]/I as an MM-graded algebra.

Let BB be a 𝐤\mathbf{k}-algebra. A derivation \partial of BB, is a 𝐤\mathbf{k}-linear map satisfying the Leibniz rule, that is,

(fg)=(f)g+f(g),for all f,gB.\partial(fg)=\partial(f)g+f\partial(g),\quad\text{for all }f,g\in B.

We denote by Der(B)\operatorname{Der}(B) the 𝐤\mathbf{k}-vector space of derivations of BB. Let IBI\subset B be an ideal. We denote by DerI(B)\operatorname{Der}_{I}(B) the subspace of derivations Der(B)\partial\in\operatorname{Der}(B) such that (I)I\partial(I)\subset I. A derivation \partial is locally nilpotent if for each fBf\in B there exists >0\ell\in\mathbb{Z}_{>0} such that (f)=0\partial^{\ell}(f)=0, where \partial^{\ell} is the composition of \partial \ell-times.

Letting SS be a semigroup, we let now \partial be a derivation of 𝐤[S]\mathbf{k}[S]. We say that \partial is homogeneous if it sends homogeneous elements into homogeneous elements with respect to the SS-grading of 𝐤[S]\mathbf{k}[S]. By [DL24, Lemma 1.1], there exists a unique element αM\alpha\in M, called the degree of \partial and denoted by deg\deg\partial, such that for every 𝐱mker\mathbf{x}^{m}\notin\ker\partial we have (𝐱m)=λ𝐱m+α\partial(\mathbf{x}^{m})=\lambda\mathbf{x}^{m+\alpha} for some λ𝐤\lambda\in\mathbf{k}. Furthermore, by [DL24, Proposition 2.1], every derivation in Der(𝐤[S])\operatorname{Der}(\mathbf{k}[S]) admits a finite decomposition =α\partial=\sum\partial_{\alpha}, where each α\partial_{\alpha} is a homogeneous derivation on 𝐤[S]\mathbf{k}[S] of degree αM\alpha\in M. We say that a homogeneous derivation \partial is inner if degS\deg\partial\in S and outer if degMS\deg\partial\in M\setminus S. By a straightforward application of the Leibniz rule, it follows that every homogeneous derivation on 𝐤[S]\mathbf{k}[S] is of the form

α,p:𝐤[S]𝐤[S],𝐱mp(m)𝐱m+α,\displaystyle\partial_{\alpha,p}\colon\mathbf{k}[S]\to\mathbf{k}[S],\quad\mathbf{x}^{m}\mapsto p(m)\cdot\mathbf{x}^{m+\alpha}\,, (2)

where α=degα,pM\alpha=\deg\partial_{\alpha,p}\in M and pN𝐤=N𝐤p\in N_{\mathbf{k}}=N\otimes_{\mathbb{Z}}\mathbf{k}.

Let now SS be an affine semigroup and I𝐤[S]I\subset\mathbf{k}[S] be a monomial ideal. The Lie algebras Der(𝐤[S])\operatorname{Der}(\mathbf{k}[S]) and DerI(𝐤[S])\operatorname{Der}_{I}(\mathbf{k}[S]) have been studied by several authors, see for instance [BS95, Tad09, Lie10, KLL15, DL24]. We present now the description of derivations of 𝐤[S]\mathbf{k}[S] given in [KLL15, Proposition 3.1] with the notation adapted to our context. To describe inner derivations, we now let

𝔤=αS𝔤αDer(𝐤[S])where𝔤α={α,ppN𝐤}.\mathfrak{g}=\bigoplus_{\alpha\in S}\mathfrak{g}_{\alpha}\subset\operatorname{Der}(\mathbf{k}[S])\quad\mbox{where}\quad\mathfrak{g}_{\alpha}=\left\{\partial_{\alpha,p}\mid p\in N_{\mathbf{k}}\right\}.

For every αS\alpha\in S, we have that 𝔤α\mathfrak{g}_{\alpha} is a vector space isomorphic to N𝐤,N_{\mathbf{k}}, since α,p+α,q=α,p+q,\partial_{\alpha,p}+\partial_{\alpha,q}=\partial_{\alpha,p+q}, for all p,qN𝐤p,q\in N_{\mathbf{k}}.

To describe outer derivations, which correspond exactly to the homogeneous locally nilpotent derivations in the notation of [KLL15] we need to introduce some notation. For every semigroup SS, we define its dual semigroup as

S={pNp(m)0, for all mS}.S^{\vee}=\{p\in N\mid p(m)\geq 0,\mbox{ for all }m\in S\}\,.

We observe that SS^{\vee} is an affine saturated semigroup, see [CLS11, Proposition 1.2.4]. For every ρS(1)\rho\in S^{\vee}(1) we define the set

=ρS(1)ρwhereρ={αMρ(α)=1 and ρ(α)0 for all ρS(1){ρ}}.\displaystyle\mathcal{R}=\bigcup_{\rho\in S^{\vee}(1)}\mathcal{R}_{\rho}\quad\mbox{where}\quad\mathcal{R}_{\rho}=\left\{\alpha\in M\mid\rho(\alpha)=-1\mbox{ and }\rho^{\prime}(\alpha)\geq 0\mbox{ for all }\rho^{\prime}\in S^{\vee}(1)\setminus\{\rho\}\right\}\,. (3)

Elements in \mathcal{R} are called the Demazure roots of the semigroup SS. Note that by definition, the sets ρ\mathcal{R}_{\rho} are disjoint since ρ(α)=1\rho(\alpha)=-1 for one and only one ρS(1)\rho\in S^{\vee}(1). We now let

𝔰=α𝔰αwhere𝔰α={α,pp=λρ, with λ𝐤}\mathfrak{s}=\bigoplus_{\alpha\in\mathcal{R}}\mathfrak{s}_{\alpha}\quad\mbox{where}\quad\mathfrak{s}_{\alpha}=\left\{\partial_{\alpha,p}\mid p=\lambda\cdot\rho,\mbox{ with }\lambda\in\mathbf{k}\right\}

is the 11-dimensional vector space of outer derivations of degree α\alpha.

Proposition 1.2 ([KLL15, Proposition 3.1]).

Let SS be a saturated affine semigroup. Then

Der(𝐤[S])=𝔤𝔰,\operatorname{Der}(\mathbf{k}[S])=\mathfrak{g}\oplus\mathfrak{s}\,,

where homogeneous derivations in 𝔤\mathfrak{g} are inner and homogeneous derivations in 𝔰\mathfrak{s} are outer.

We now provide a description of DerI(𝐤[S])\operatorname{Der}_{I}(\mathbf{k}[S]). The image (I)\partial(I) is contained in II for all inner derivations 𝔤\partial\in\mathfrak{g}. So 𝔤DerI(𝐤[S])\mathfrak{g}\subset\operatorname{Der}_{I}(\mathbf{k}[S]). For outer derivations, let 𝔰(I)={𝔰(I)I}\mathfrak{s}(I)=\{\partial\in\mathfrak{s}\mid\partial(I)\subset I\}. Letting I𝐤[S]I\subset\mathbf{k}[S] be the ideal I=(𝐱𝐚1,,𝐱𝐚l)I=(\mathbf{x}^{\mathbf{a}_{1}},\dots,\mathbf{x}^{\mathbf{a}_{l}}) with 𝐚iS\mathbf{a}_{i}\in S, we let

(I)=ρS(1)ρ(I)whereρ(I)={αρ𝐚i+αsupp(I) for all 𝐚iρ}.\mathcal{R}(I)=\bigcup_{\rho\in S^{\vee}(1)}\mathcal{R}_{\rho}(I)\quad\mbox{where}\quad\mathcal{R}_{\rho}(I)=\left\{\alpha\in\mathcal{R}_{\rho}\mid\mathbf{a}_{i}+\alpha\in\operatorname{supp}(I)\mbox{ for all }\mathbf{a}_{i}\notin\rho^{\bot}\right\}\,.

The following proposition provides a description of DerI(𝐤[S])\operatorname{Der}_{I}(\mathbf{k}[S]).

Proposition 1.3.

Let SS be a saturated affine semigroup and let I=(𝐱𝐚1,,𝐱𝐚l)I=(\mathbf{x}^{\mathbf{a}_{1}},\dots,\mathbf{x}^{\mathbf{a}_{l}}) with 𝐚iS\mathbf{a}_{i}\in S. Then

𝔰(I)=α(I)𝔰α.\mathfrak{s}(I)=\bigoplus_{\alpha\in\mathcal{R}(I)}\mathfrak{s}_{\alpha}.

In particular DerI(𝐤[S])=𝔤𝔰(I).\operatorname{Der}_{I}(\mathbf{k}[S])=\mathfrak{g}\oplus\mathfrak{s}(I).

Proof.

Let 𝔰(I).\partial\in\mathfrak{s}(I). Then there exists r>0r\in\mathbb{Z}_{>0} such that =i=1rαi,\partial=\sum_{i=1}^{r}\partial_{\alpha_{i}}, where αi=degαi,αi𝔰αi\alpha_{i}=\deg\partial_{\alpha_{i}},\>\partial_{\alpha_{i}}\in\mathfrak{s}_{\alpha_{i}} and αi\partial_{\alpha_{i}} is homogeneous for all i{1,,r}.i\in\{1,\ldots,r\}. Let α=αi,\alpha=\alpha_{i}, for some i{1,,r},i\in\{1,\ldots,r\}, and let ρN\rho\in N be such that αρ.\alpha\in\mathcal{R}_{\rho}. We claim that αρ(I).\alpha\in\mathcal{R}_{\rho}(I). Let 𝐚{𝐚1,,𝐚l}ρ.\mathbf{a}\in\{\mathbf{a}_{1},\ldots,\mathbf{a}_{l}\}\setminus\rho^{\perp}. Then (𝐱𝐚)=αi(𝐱𝐚)I.\partial(\mathbf{x}^{\mathbf{a}})=\sum\partial_{\alpha_{i}}(\mathbf{x}^{\mathbf{a}})\in I. It follows by [HH11, Corollary 1.1.3] that α(𝐱𝐚)I,\partial_{\alpha}(\mathbf{x}^{\mathbf{a}})\in I, that is 𝐚+αsupp(I).\mathbf{a}+\alpha\in\operatorname{supp}(I). Hence, αρ(I).\alpha\in\mathcal{R}_{\rho}(I). The second statement follows directly from Proposition 1.2. ∎

The above description of 𝔰(I)\mathfrak{s}(I) was given in [Tad09, Theorem 2.2] for the case where 𝐤[S]\mathbf{k}[S] is the polynomial ring, as in Example 1.1, see also [BS95, Theorem 2.2.1]. We apply Proposition 1.3 to the polynomial ring in the following example.

Example 1.4 (Derivations of the polynomial ring).

With the notation of Example 1.1, we have that the inner derivations of the polynomial ring are of the form

α,p=x1α1xnαn(p1x1ddx1++pnxnddxn),\partial_{\alpha,p}=x_{1}^{\alpha_{1}}\ldots x_{n}^{\alpha_{n}}\cdot\left(p_{1}x_{1}\frac{d}{dx_{1}}+\cdots+p_{n}x_{n}\frac{d}{dx_{n}}\right),

for α=(α1,,αn)0n\alpha=(\alpha_{1},\ldots,\alpha_{n})\in\mathbb{Z}_{\geq 0}^{n} and p=(p1,,pn)𝐤n.p=(p_{1},\dots,p_{n})\in\mathbf{k}^{n}. We now compute the outer derivations. Let E={e1,,en}E^{*}=\{e_{1}^{*},\ldots,e_{n}^{*}\} be the canonical basis of N.N. Then EE^{*} is the set of generating vectors of the rays of S.S^{*}. Thus, for 1jn,1\leq j\leq n, we have

ej={α=(α1,,αn)nαj=1 and αi0 for all ij}.\mathcal{R}_{e_{j}^{*}}=\big{\{}\alpha=(\alpha_{1},\ldots,\alpha_{n})\in\mathbb{Z}^{n}\mid\alpha_{j}=-1\mbox{ and }\alpha_{i}\geq 0\mbox{ for all }i\neq j\big{\}}.

Hence, =j=1nej\mathcal{R}=\bigcup_{j=1}^{n}\mathcal{R}_{e_{j}^{*}} yields the degree of all the outer derivations on 𝐤[S].\mathbf{k}[S]. Thus, for 1jn,1\leq j\leq n, the derivation α,p\partial_{\alpha,p} where p=λej,p=\lambda\cdot e_{j}^{*}, for some λ𝐤\lambda\in\mathbf{k}^{*} and α=(α1,,αn)ej,\alpha=(\alpha_{1},\ldots,\alpha_{n})\in\mathcal{R}_{e_{j}^{*}}, corresponds to

α,p=λx1α1xjαj^xnαnddxj,\partial_{\alpha,p}=\lambda\cdot x_{1}^{\alpha_{1}}\cdots\widehat{x_{j}^{\alpha_{j}}}\cdots x_{n}^{\alpha_{n}}\cdot\frac{d}{dx_{j}}\,,

where xjαj^\widehat{x_{j}^{\alpha_{j}}} means that the factor is excluded from the product.

Example 1.5.

With the notation of Example 1.1, we let n=2n=2. Let I𝐤[x,y]I\subset\mathbf{k}[x,y] be the monomial ideal (x2y5,x3y2,x5)(x^{2}y^{5},x^{3}y^{2},x^{5}). We compute (I).\mathcal{R}(I). The support supp(I)\operatorname{supp}(I) is presented in Figure 4. Let 𝐚1=(2,5),𝐚2=(3,2),𝐚3=(5,0)02\mathbf{a}_{1}=(2,5),\mathbf{a}_{2}=(3,2),\mathbf{a}_{3}=(5,0)\in\mathbb{Z}_{\geq 0}^{2} be the exponents of the generators of I.I. We first note that e1\mathcal{R}_{e_{1}^{*}} is empty, because 𝐚1+α=(1,+5)supp(I),\mathbf{a}_{1}+\alpha=(1,\ell+5)\notin\operatorname{supp}(I), for any α=(1,),\alpha=(-1,\ell), where 0.\ell\in\mathbb{Z}_{\geq 0}.

We now describe e2(I).\mathcal{R}_{e_{2}^{*}}(I). Note that e2\mathcal{R}_{e_{2}^{*}} is the set of elements α=(,1),\alpha=(\ell,-1), where 0,\ell\in\mathbb{Z}_{\geq 0}, such that α+𝐚1=(2+,4),α+𝐚2=(3+,1)supp(I),\alpha+\mathbf{a}_{1}=(2+\ell,4),\alpha+\mathbf{a}_{2}=(3+\ell,1)\in\operatorname{supp}(I), because e2(𝐚3)=0.e_{2}^{*}(\mathbf{a}_{3})=0. Thus e2(I)={(,1)22}.\mathcal{R}_{e_{2}^{*}}(I)=\{(\ell,-1)\in\mathbb{Z}^{2}\mid\ell\geq 2\}. Hence, we have

𝔰(I)=αe2(I)𝔰α==2𝐤xddy=(=2𝐤x)ddy=x2𝐤[x]ddy.\mathfrak{s}(I)=\bigoplus_{\alpha\in\mathcal{R}_{e_{2}^{*}}(I)}\mathfrak{s}_{\alpha}=\bigoplus_{\ell=2}^{\infty}\mathbf{k}\cdot x^{\ell}\frac{d}{dy}=\left(\bigoplus_{\ell=2}^{\infty}\mathbf{k}\cdot x^{\ell}\right)\frac{d}{dy}=x^{2}\mathbf{k}[x]\cdot\frac{d}{dy}.

The set (I)=e2(I)\mathcal{R}(I)=\mathcal{R}_{e_{2}^{*}}(I) is given by the red dots in in Figure 4 represents the degrees of outer derivations DerI(𝐤[x,y])\partial\in\operatorname{Der}_{I}(\mathbf{k}[x,y]). Finally, we observe that all red dots represent trivial derivations on 𝐤[x,y]/I,\mathbf{k}[x,y]/I, except those of degree α{(2,1),(3,1),(4,1)}\alpha\in\{(2,-1),(3,-1),(4,-1)\}.

Figure 3. I=(x2y5,x3y2,x5)I=(x^{2}y^{5},x^{3}y^{2},x^{5})
Figure 4. (I)={(,1)2}\mathcal{R}(I)=\{(\ell,-1)\mid\ell\geq 2\}

2. Derivations and Monomial Ideals

Let SS be an affine semigroup and I𝐤[S]I\subset\mathbf{k}[S] a monomial ideal. Consider the natural projection map

π:DerI(𝐤[S])Der(𝐤[S]/I),¯defined by¯(a¯)=(a)¯.\displaystyle\pi\colon\operatorname{Der}_{I}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I),\quad\partial\mapsto\overline{\partial}\quad\text{defined by}\quad\overline{\partial}(\overline{a})=\overline{\partial(a)}. (4)

This map π\pi is not always surjective. Indeed, the following example, provided by Kraft in his unpublished notes [Kra17, Example 5], demonstrates non-surjectivity in a specific case.

Example 2.1.

Let M=2M=\mathbb{Z}^{2} and let SS be the semigroup generated by μ1=(1,0)\mu_{1}=(1,0), μ2=(1,1)\mu_{2}=(1,1), and μ3=(0,2)\mu_{3}=(0,2). Note that SS is not saturated because (0,1)ωSM(0,1)\in\omega_{S}\cap M, but (0,1)S(0,1)\notin S. Furthermore, X=Spec𝐤[S]X=\operatorname{Spec}\mathbf{k}[S] is isomorphic to the variety in the affine space 𝔸3\mathbb{A}^{3} defined by the equation x2zy2=0x^{2}z-y^{2}=0, by setting x=𝐱μ1x=\mathbf{x}^{\mu_{1}}, y=𝐱μ2y=\mathbf{x}^{\mu_{2}}, and z=𝐱μ3z=\mathbf{x}^{\mu_{3}} (commonly known as Whitney’s Umbrella). It follows from [DL24, Theorem 3.11] that ={(,1)21}\mathcal{R}=\{(\ell,-1)\in\mathbb{Z}^{2}\mid\ell\geq 1\}. This is represented by the green points in Figure 6. Let I𝐤[S]I\subset\mathbf{k}[S] be the monomial ideal generated by 𝐚1=(1,0)\mathbf{a}_{1}=(1,0) and 𝐚2=(1,1)\mathbf{a}_{2}=(1,1), shown in Figure 6. The ideal II corresponds to I=(x,y)I=(x,y) under the isomorphism mentioned above, and 𝐤[S]/I𝐤[z¯]\mathbf{k}[S]/I\simeq\mathbf{k}[\overline{z}]. Since ddz¯\dfrac{d}{d\overline{z}} is a homogeneous derivation on 𝐤[z¯]\mathbf{k}[\overline{z}] of degree (0,2)(0,-2), we conclude that π\pi cannot be surjective because (0,2)(0,-2)\notin\mathcal{R}.

Figure 5. S={(0,2),(1,1),(1,0)}0S=\{(0,2),(1,1),(1,0)\}\mathbb{Z}_{\geq 0}
Figure 6. I=(x,y)I=(x,y)

Inspired by the above example, which concerns a non-saturated affine semigroup, we present the following theorem, which is the aim of this section. In this theorem, we show that non-surjectivity can occur for any saturated affine semigroup that is not isomorphic to the first octant 0n\mathbb{Z}^{n}_{\geq 0}. Recall that, unless stated otherwise, all our semigroups are affine, saturated, pointed and minimally embedded in MM.

Theorem 2.2.

Let SS be a semigroup. The natural map

π:DerI(𝐤[S])Der(𝐤[S]/I)\pi\colon\operatorname{Der}_{I}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I)

is surjective for every monomial ideal II if and only if S0nS\simeq\mathbb{Z}^{n}_{\geq 0}.

We divide the proof into several lemmas. Let SS be a saturated affine semigroup. We say that mS{0}m\in S\setminus\{0\} is irreducible if m=m1+m2m=m_{1}+m_{2} implies m1=0m_{1}=0 or m2=0m_{2}=0. The set \mathcal{H} of all irreducible elements is called the Hilbert basis of SS. Since SS is pointed, the Hilbert basis \mathcal{H} is the unique minimal generating set of SS as a semigroup. Let ={μ1,,μl}S\mathcal{H}=\{\mu_{1},\dots,\mu_{l}\}\subset S be the Hilbert basis of SS. In all cases where SS is different from the first octant 0n\mathbb{Z}_{\geq 0}^{n}, we will prove that there exists a non-liftable derivation in 𝐤[S]/I\mathbf{k}[S]/I_{\mathcal{H}}, where II_{\mathcal{H}} is the monomial ideal

I=mS𝐤𝐱m.I_{\mathcal{H}}=\bigoplus_{m\in S\setminus\mathcal{H}}\mathbf{k}\cdot\mathbf{x}^{m}\,.

Let ρS(1)\rho\in S^{\vee}(1). Note that there exist m1,m2m_{1},m_{2}\in\mathcal{H} such that ρ(m1)=0\rho(m_{1})=0 and ρ(m2)=1\rho(m_{2})=1, since \mathcal{H} generates SS.

Lemma 2.3.

Let SS be a semigroup. If there exists ρS(1)\rho\in S^{\vee}(1) such that {0,1}\{0,1\} is a proper subset of ρ()\rho(\mathcal{H}), then the natural map π:DerI(𝐤[S])Der(𝐤[S]/I)\pi\colon\operatorname{Der}_{I_{\mathcal{H}}}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I_{\mathcal{H}}) is not surjective.

Proof.

Let ={μ1,,μl}\mathcal{H}=\{\mu_{1},\dots,\mu_{l}\} be the Hilbert basis of SS. By hypothesis, there exists ρS(1)\rho\in S^{*}(1) such that the set {0,1}\{0,1\} is a proper subset of ρ().\rho(\mathcal{H}). Without loss of generality, we assume that ρ(μ1)=>1\rho(\mu_{1})=\ell>1 and ρ(μ2)=0\rho(\mu_{2})=0. We define the linear map

¯:𝐤[S]/I𝐤[S]/I,given by¯(𝐱¯μ1)=𝐱¯μ2 and ¯(𝐱¯μi)=0 for all i>1.\overline{\partial}\colon\mathbf{k}[S]/I_{\mathcal{H}}\to\mathbf{k}[S]/I_{\mathcal{H}},\quad\mbox{given by}\quad\overline{\partial}(\overline{\mathbf{x}}^{\mu_{1}})=\overline{\mathbf{x}}^{\mu_{2}}\mbox{ and }\overline{\partial}(\overline{\mathbf{x}}^{\mu_{i}})=0\mbox{ for all }i>1\,.

The map ¯\overline{\partial} is well defined since ¯(I)=0\overline{\partial}(I)=0. Moreover, it is a derivation. Indeed, the only non-trivial instance of the Leibniz rule is

0=¯(𝐱¯μ1+μi)=¯(𝐱¯μ1)𝐱¯μi+𝐱¯μ1¯(𝐱¯μi)=𝐱¯μ1¯(𝐱¯μi).0=\overline{\partial}(\overline{\mathbf{x}}^{\mu_{1}+\mu_{i}})=\overline{\partial}(\overline{\mathbf{x}}^{\mu_{1}})\overline{\mathbf{x}}^{\mu_{i}}+\overline{\mathbf{x}}^{\mu_{1}}\overline{\partial}(\overline{\mathbf{x}}^{\mu_{i}})=\overline{\mathbf{x}}^{\mu_{1}}\overline{\partial}(\overline{\mathbf{x}}^{\mu_{i}})\,.

Note that ¯(𝐱¯μi)\overline{\partial}(\overline{\mathbf{x}}^{\mu_{i}}) is equal to 0 or 𝐱¯μ2\overline{\mathbf{x}}^{\mu_{2}}. In both cases, the Leibniz rule is verified since

𝐱¯μ1¯(𝐱¯μi)=0.\overline{\mathbf{x}}^{\mu_{1}}\overline{\partial}(\overline{\mathbf{x}}^{\mu_{i}})=0\,.

The derivation ¯\overline{\partial} is homogeneous and its degree is deg¯=α=μ2μ1\deg\overline{\partial}=\alpha=\mu_{2}-\mu_{1} and ρ(α)=<1\rho(\alpha)=-\ell<-1. Assume that there exists αDer(𝐤[S])\partial_{\alpha}\in\operatorname{Der}(\mathbf{k}[S]) homogeneous such that π(α)=¯α\pi(\partial_{\alpha})=\overline{\partial}_{\alpha}. Since ρ(α)\rho(\alpha) is negative, α\partial_{\alpha} must be an outer derivation. Finally, for all outer derivations we have ρ(α)1\rho(\alpha)\geq-1 by (3). This provides a contradiction. Hence, the lifting α\partial_{\alpha} does not exist. ∎

In the proof of Lemma 2.6, we need the following technical lemma.

Lemma 2.4.

Let MM be a free abelian group of rank nn and let NN be the dual group N=Hom(M,)N=\operatorname{Hom}(M,\mathbb{Z}). Let β={μ1,,μn}M\beta=\{\mu_{1},\ldots,\mu_{n}\}\subset M and β={ρ1,,ρn}N\beta^{\prime}=\{\rho_{1},\ldots,\rho_{n}\}\subset N be linearly independent sets. We let MM^{\prime} be the subgroup of MM generated by β\beta and NN^{\prime} be the subgroup of NN generated by β\beta^{\prime}. If A=(aij)A=(a_{ij}) is the n×nn\times n matrix with aij=ρi(μj)a_{ij}=\rho_{i}(\mu_{j}), then

|M/M||N/N|=|det(A)|.|M/M^{\prime}|\cdot|N/N^{\prime}|=|\det(A)|\,.
Proof.

The lemma follows directly by taking dual bases β\beta and β\beta^{*} of MM and NN, respectively and applying [Bar02, Chapter VII, Section 2, Theorem 2.5] to the subgroups MMM^{\prime}\subset M and NNN^{\prime}\subset N on both sides of the duality. ∎

To prove Theorem 2.2, we introduce the following notation that will be convenient in the proof.

Definition 2.5.

Let SS be a semigroup minimally embedded in MM with rankM=n\operatorname{rank}M=n. We define a complete flag \mathcal{F} of SS as a chain of faces FiSF_{i}\subseteq S such that

{0}=F0F1Fn=SwithrankFi=i.\{0\}=F_{0}\subset F_{1}\subset\ldots\subset F_{n}=S\quad\text{with}\quad\operatorname{rank}F_{i}=i\,.

The complete flag \mathcal{F} also defines a chain of reverse inclusions

S=F0F1Fn={0}withrankFi=ni,S^{\vee}=F^{*}_{0}\supset F^{*}_{1}\supset\ldots\supset F^{*}_{n}=\{0\}\quad\text{with}\quad\operatorname{rank}F^{*}_{i}=n-i\,,

where FiF^{*}_{i} is the face of SNS^{\vee}\subset N dual to FiF_{i}, denoted as Fi=SFiF^{*}_{i}=S^{\vee}\cap F_{i}^{\perp}.

Given a complete flag, an upper triangular pair of rays is a pair of sets β={μ1,,μn}S(1)\beta=\{\mu_{1},\ldots,\mu_{n}\}\subset S(1) and β={ρ1,,ρn}S(1)\beta^{\prime}=\{\rho_{1},\ldots,\rho_{n}\}\subset S^{\vee}(1) such that

μiFiFi1andρiFi1Fi,\mu_{i}\in F_{i}\setminus F_{i-1}\quad\text{and}\quad\rho_{i}\in F^{*}_{i-1}\setminus F^{*}_{i}\,,

satisfying

ρi(μj)={0ifi>jifijandρi(μi)>0, for all i.\displaystyle\rho_{i}(\mu_{j})=\begin{cases}0&\text{if}\quad i>j\\ *&\text{if}\quad i\leq j\end{cases}\quad\text{and}\quad\rho_{i}(\mu_{i})>0,\text{ for all }i\,. (5)

In the following lemma, we show that an upper triangular pair always exists.

Lemma 2.6.

Let SS be a semigroup and let \mathcal{H} be its Hilbert basis. Let \mathcal{F} be a complete flag of SS. Then, there exists an upper triangular pair (β,β)(\beta,\beta^{\prime}). Moreover, if ρ(S(1))={0,1}\rho(S(1))=\{0,1\} for all ρS(1)\rho\in S^{\vee}(1), then β\beta and β\beta^{\prime} are \mathbb{Z}-bases of MM and NN, respectively.

Proof.

Let SS be minimally embedded in MM with rankM=n\operatorname{rank}M=n. The complete flag \mathcal{F} provides us with the inclusions of faces

{0}=F0F1Fn=SandS=F0F1Fn={0}.\{0\}=F_{0}\subset F_{1}\subset\ldots\subset F_{n}=S\quad\text{and}\quad S^{\vee}=F^{*}_{0}\supset F^{*}_{1}\supset\ldots\supset F^{*}_{n}=\{0\}\,.

We will appropriately choose the sets β={μ1,,μn}S(1)\beta=\{\mu_{1},\ldots,\mu_{n}\}\subset S(1) and β={ρ1,,ρn}S(1)\beta^{\prime}=\{\rho_{1},\ldots,\rho_{n}\}\subset S^{\vee}(1) to obtain an upper triangular pair (β,β)(\beta,\beta^{\prime}). First, we pick μi\mu_{i} as any ray in FiFi1F_{i}\setminus F_{i-1}. Then, we choose ρi\rho_{i} to be any ray in Fi1F^{*}_{i-1} with ρi(μi)>0\rho_{i}(\mu_{i})>0. Such a ray must exist; otherwise, μiFi1\mu_{i}\in F_{i-1}, which is a contradiction. This choice satisfies the condition in (5). Indeed, if i>ji>j, then ρi\rho_{i} is a ray in Fi1FjF^{*}_{i-1}\subseteq F^{*}_{j} and μj\mu_{j} is a ray in FjF_{j}, so ρi(μj)=0\rho_{i}(\mu_{j})=0. This proves the first statement of the lemma.

To prove the second statement, we assume, moreover, that ρ(S(1))={0,1}\rho(S(1))=\{0,1\}. Under these conditions, ρi(μi)>0\rho_{i}(\mu_{i})>0 in (5) implies ρi(μi)=1\rho_{i}(\mu_{i})=1. Let A=(aij)A=(a_{ij}) be the n×nn\times n matrix with aij=ρi(μj)a_{ij}=\rho_{i}(\mu_{j}). By (5) and the fact that ρi(μi)=1\rho_{i}(\mu_{i})=1, we have that det(A)=1\det(A)=1, so β\beta and β\beta^{\prime} are linearly independent sets. Moreover, by Lemma 2.4, we have that |M/M|=|N/N|=1|M/M^{\prime}|=|N/N^{\prime}|=1, where MM^{\prime} and NN^{\prime} are the submodules spanned by β\beta and β\beta^{\prime} in MM and NN, respectively. We conclude that β\beta and β\beta^{\prime} are \mathbb{Z}-bases of MM and NN, respectively. ∎

Corollary 2.7.

Let SS be a simplicial affine semigroup. If SS is not isomorphic to 0n\mathbb{Z}_{\geq 0}^{n}, then the natural map π:DerI(𝐤[S])Der(𝐤[S]/I)\pi\colon\operatorname{Der}_{I_{\mathcal{H}}}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I_{\mathcal{H}}) is not surjective.

Proof.

If SS is not isomorphic to 0n\mathbb{Z}_{\geq 0}^{n}, then the set S(1)S(1) is not a \mathbb{Z}-basis of MM. Now, take any complete flag of SS. According to Lemma 2.6, we can order the sets S(1)S(1) and S(1)S^{\vee}(1), both with nn elements, to form an upper triangular pair. Furthermore, by the second statement in Lemma 2.6, there exists ρS(1)\rho\in S^{\vee}(1) such that {0,1}\{0,1\} is a proper subset of ρ(S(1))ρ()\rho(S(1))\subset\rho(\mathcal{H}), since S(1)S(1) is not a \mathbb{Z}-basis of MM. Thus, the non-surjectivity of π\pi follows from Lemma 2.3. ∎

In the following lemma, we handle the case where SS is not simplicial.

Lemma 2.8.

Let SS be a semigroup. If SS is not simplicial, then the natural map π:DerI(𝐤[S])Der(𝐤[S]/I)\pi\colon\operatorname{Der}_{I_{\mathcal{H}}}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I_{\mathcal{H}}) is not surjective.

Proof.

Let ={μ1,,μl}\mathcal{H}=\{\mu_{1},\ldots,\mu_{l}\} be the Hilbert basis of SS and let \mathcal{F} be a complete flag that gives us the inclusions of faces

{0}=F0F1Fn=SandS=F0F1Fn={0}.\{0\}=F_{0}\subset F_{1}\subset\ldots\subset F_{n}=S\quad\text{and}\quad S^{\vee}=F^{*}_{0}\supset F^{*}_{1}\supset\ldots\supset F^{*}_{n}=\{0\}\,.

By Lemma 2.6, there exists an upper triangular pair (β,β)(\beta,\beta^{\prime}). Recall that

μiFiFi1andρiFi1Fi.\mu_{i}\in F_{i}\setminus F_{i-1}\quad\text{and}\quad\rho_{i}\in F^{*}_{i-1}\setminus F^{*}_{i}\,.

Now we have that Fn1=ρnSF_{n-1}=\rho_{n}^{\perp}\cap S and G=ρn1SG=\rho_{n-1}^{\perp}\cap S are facets whose intersection

Fn1G=ρn1ρnS=Fn2F_{n-1}\cap G=\rho_{n-1}^{\perp}\cap\rho_{n}^{\perp}\cap S=F_{n-2}

is an (n2)(n-2)-dimensional face. Since SS is not simplicial, up to changing the complete flag \mathcal{F}, we can assume that there exists another ray μS(1)\mu\in S(1) that is not contained in Fn1GF_{n-1}\cup G. Hence, ρn(μ)>0\rho_{n}(\mu)>0 and ρn1(μ)>0\rho_{n-1}(\mu)>0. We now define the linear map

¯:𝐤[S]/I𝐤[S]/I,given by¯(𝐱¯μ)=𝐱¯μ1 and ¯(𝐱¯μ)=0 for all μ{μ},\overline{\partial}\colon\mathbf{k}[S]/I_{\mathcal{H}}\to\mathbf{k}[S]/I_{\mathcal{H}},\quad\text{given by}\quad\overline{\partial}(\overline{\mathbf{x}}^{\mu})=\overline{\mathbf{x}}^{\mu_{1}}\text{ and }\overline{\partial}(\overline{\mathbf{x}}^{\mu^{\prime}})=0\text{ for all }\mu^{\prime}\in\mathcal{H}\setminus\{\mu\}\,,

where μ1\mu_{1} is the first element in the Hilbert basis \mathcal{H}. Using the same argument as in Lemma 2.3, we conclude that ¯\overline{\partial} is a derivation of 𝐤[S]/I\mathbf{k}[S]/I.

The degree of ¯\overline{\partial} is α=μ1μ\alpha=\mu_{1}-\mu, and so ρn1(α)<0\rho_{n-1}(\alpha)<0 and ρn(α)<0\rho_{n}(\alpha)<0. Assume that there exists Der(𝐤[S])\partial\in\operatorname{Der}(\mathbf{k}[S]) homogeneous such that π()=¯\pi(\partial)=\overline{\partial}. The degree of \partial must also be α\alpha and since ρn1(α)<0\rho_{n-1}(\alpha)<0 and ρn(α)<0\rho_{n}(\alpha)<0, the derivation α\partial_{\alpha} must be an outer derivation. Finally, by (3), for all outer derivations, we have that ρ(α)\rho(\alpha) is negative for one and only one ρS(1)\rho\in S^{\vee}(1). This provides a contradiction. ∎

We now deal with the case where 𝐤[S]\mathbf{k}[S] is the polynomial ring.

Lemma 2.9.

Let S=0nS=\mathbb{Z}_{\geq 0}^{n} and let II be a monomial ideal in 𝐤[S]\mathbf{k}[S]. Then, the natural map π:DerI(𝐤[S])Der(𝐤[S]/I)\pi\colon\operatorname{Der}_{I}(\mathbf{k}[S])\to\operatorname{Der}(\mathbf{k}[S]/I) is surjective.

Proof.

Recall that 𝐤[S]=𝐤[x1,,xn]\mathbf{k}[S]=\mathbf{k}[x_{1},\dots,x_{n}] as in Example 1.1, and let φ:𝐤[S]𝐤[S]/I\varphi\colon\mathbf{k}[S]\to\mathbf{k}[S]/I be the quotient map. For any derivation ¯Der(𝐤[S]/I)\overline{\partial}\in\operatorname{Der}(\mathbf{k}[S]/I), we choose fiφ1(¯(x¯i))f_{i}\in\varphi^{-1}(\overline{\partial}(\overline{x}_{i})). Since Der(𝐤[S])\operatorname{Der}(\mathbf{k}[S]) is freely generated by the partial derivatives {ddx1,,ddxn}\left\{\dfrac{d}{dx_{1}},\dots,\dfrac{d}{dx_{n}}\right\} as a 𝐤[S]\mathbf{k}[S]-module [Fre17, Section 3.2.2], the map :𝐤[S]𝐤[S]\partial\colon\mathbf{k}[S]\to\mathbf{k}[S] defined by

=i=1nfiddxiis a derivation of 𝐤[S] such that (xi)=fi.\partial=\sum_{i=1}^{n}f_{i}\dfrac{d}{dx_{i}}\quad\text{is a derivation of }\mathbf{k}[S]\text{ such that }\partial(x_{i})=f_{i}\,.

Since ¯φ=φ\overline{\partial}\circ\varphi=\varphi\circ\partial, we have that (I)I\partial(I)\subset I. This proves that π()=¯\pi(\partial)=\overline{\partial} and so π\pi is surjective. ∎

We can finally proceed with the proof of Theorem 2.2.

Proof of Theorem 2.2.

The direct implication follows from Corollary 2.7 if SS is simplicial and from Lemma 2.8 if SS is not simplicial. On the other hand, the converse implication follows from Lemma 2.9. ∎

3. Liftable locally nilpotent derivation of monomial algebras

Let SS be a semigroup and let I𝐤[S]I\subset\mathbf{k}[S] be a monomial ideal. In this section, we compute the homogeneous locally nilpotent derivations of the algebra 𝐤[S]/I\mathbf{k}[S]/I, which are obtained as the images of derivations of 𝐤[S]\mathbf{k}[S] via π\pi in (4). We formalize this concept in the following definition.

Definition 3.1.

Let SS be a semigroup and let II be a monomial ideal of 𝐤[S]\mathbf{k}[S]. We say that a derivation ¯\overline{\partial} on 𝐤[S]/I\mathbf{k}[S]/I is liftable if there exists a derivation DerI(𝐤[S])\partial\in\operatorname{Der}_{I}(\mathbf{k}[S]) such that ¯=π()\overline{\partial}=\pi(\partial).

Remark 3.2.

In view of Theorem 2.2, if SS is isomorphic to 0n\mathbb{Z}_{\geq 0}^{n}, then all derivations of 𝐤[S]/I\mathbf{k}[S]/I are liftable.

If ¯:𝐤[S]/I𝐤[S]/I\overline{\partial}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I is a liftable homogeneous derivation, then there exists a homogeneous derivation \partial of 𝐤[S]\mathbf{k}[S] such that ¯=π()\overline{\partial}=\pi(\partial). According to Proposition 1.3, we have =α,p\partial=\partial_{\alpha,p}. Therefore, we conclude that ¯=π(α,p)\overline{\partial}=\pi(\partial_{\alpha,p}), and we denote this derivation by ¯α,p\overline{\partial}_{\alpha,p}. Homogeneous derivations in 𝐤[S]\mathbf{k}[S] come in two types: inner and outer. The case of outer derivations of 𝐤[S]\mathbf{k}[S] is straightforward since they are all locally nilpotent, and any locally nilpotent derivation that descends to the quotient 𝐤[S]/I\mathbf{k}[S]/I remains locally nilpotent. We formalize this in the following lemma for future reference.

Lemma 3.3.

Let SS be a saturated affine semigroup, and let I𝐤[S]I\subset\mathbf{k}[S] be a monomial ideal. Then, any homogeneous derivation 𝔰(I)\partial\in\mathfrak{s}(I) induces a locally nilpotent derivation of 𝐤[S]/I\mathbf{k}[S]/I.

We now turn our attention to inner derivations. Inner derivations are never locally nilpotent on 𝐤[S]\mathbf{k}[S], but they may become so in the quotient 𝐤[S]/I\mathbf{k}[S]/I. Recall that all outer derivations of 𝐤[S]\mathbf{k}[S] leave II invariant and, therefore, pass to the quotient 𝐤[S]/I\mathbf{k}[S]/I.

Lemma 3.4.

Let II be a monomial ideal of 𝐤[S]\mathbf{k}[S] and let αS\alpha\in S. If (0α)supp(I)(\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I)\neq\emptyset, then the derivation ¯α,p:𝐤[S]/I𝐤[S]/I{\overline{\partial}}_{\alpha,p}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I is locally nilpotent for all pN𝐤p\in N_{\mathbf{k}}.

Proof.

By hypothesis, there exists 0\ell\in\mathbb{Z}_{\geq 0} such that αsupp(I)\ell\cdot\alpha\in\operatorname{supp}(I). Let mSm\in S. A straightforward computation shows that

α,p(𝐱m)=r𝐱m+αwherer=j=1(p(m)+(j1)p(α)).\displaystyle\partial_{\alpha,p}^{\ell}(\mathbf{x}^{m})=r\cdot\mathbf{x}^{m+\ell\cdot\alpha}\quad\mbox{where}\quad r=\prod_{j=1}^{\ell}\big{(}p(m)+(j-1)p(\alpha)\big{)}\,. (6)

Now, α,p(𝐱m)=r𝐱m+α=r𝐱m𝐱α\partial_{\alpha,p}^{\ell}(\mathbf{x}^{m})=r\cdot\mathbf{x}^{m+\ell\cdot\alpha}=r\cdot\mathbf{x}^{m}\cdot\mathbf{x}^{\ell\cdot\alpha} and 𝐱αI\mathbf{x}^{\ell\cdot\alpha}\in I since αsupp(I)\ell\cdot\alpha\in\operatorname{supp}(I). This yields α,p(𝐱m)I\partial_{\alpha,p}^{\ell}(\mathbf{x}^{m})\in I and so ¯α,p(𝐱m)=0{\overline{\partial}}_{\alpha,p}^{\ell}(\mathbf{x}^{m})=0, proving that ¯α,p(𝐱m){\overline{\partial}}_{\alpha,p}^{\ell}(\mathbf{x}^{m}) is locally nilpotent. ∎

Lemma 3.5.

Let II be a monomial ideal on 𝐤[S]\mathbf{k}[S] and let αS\alpha\in S. Assume that (0α)supp(I)=(\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I)=\emptyset. Then the derivation ¯α,p:𝐤[S]/I𝐤[S]/I{\overline{\partial}}_{\alpha,p}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I is locally nilpotent if and only if

(m+0α)supp(I)for allmSp.\displaystyle(m+\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I)\neq\emptyset\quad\mbox{for all}\quad m\in S\setminus p^{\bot}\,. (7)
Proof.

We claim that both propositions of the “if and only if” imply that p(α)=0p(\alpha)=0. Indeed, condition (7) applied to α\alpha implies αp\alpha\in p^{\bot}. On the other hand, if ¯α,p{\overline{\partial}}_{\alpha,p} is locally nilpotent, then α,p(𝐱α)=!p(α)𝐱(+1)α\partial_{\alpha,p}^{\ell}(\mathbf{x}^{\alpha})=\ell!p(\alpha)^{\ell}\mathbf{x}^{(\ell+1)\cdot\alpha} by (6) and this expression belongs to the ideal only if p(α)=0p(\alpha)=0.

Now, since p(α)=0p(\alpha)=0, we have that

α,p(𝐱m)=p(m)𝐱m+α.\displaystyle\partial_{\alpha,p}^{\ell}(\mathbf{x}^{m})=p(m)^{\ell}\cdot\mathbf{x}^{m+\ell\cdot\alpha}\,. (8)

The derivation ¯α,p{\overline{\partial}}_{\alpha,p} is locally nilpotent if and only if α,p(𝐱m)I\partial_{\alpha,p}^{\ell}(\mathbf{x}^{m})\in I for all mSm\in S. By (8) and the same argument in Lemma 3.4, this is the case if and only if

m+αsupp(I) for some 0orp(m)=0,\displaystyle m+\ell\cdot\alpha\in\operatorname{supp}(I)\mbox{ for some }\ell\in\mathbb{Z}_{\geq 0}\quad\mbox{or}\quad p(m)=0\,,

and this last condition is equivalent to (7). ∎

Remark 3.6.

It is sufficient to verify the condition in Lemma 3.5 on a set of generators of SS.

We now state our main classification result for liftable homogeneous locally nilpotent derivations on 𝐤[S]/I\mathbf{k}[S]/I. It amounts to the recollection of the three previous lemmas.

Theorem 3.7.

Let SS be a semigroup, and let II be a monomial ideal on 𝐤[S]\mathbf{k}[S]. A liftable homogeneous derivation ¯α,p:𝐤[S]/I𝐤[S]/I\overline{\partial}_{\alpha,p}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I with pN𝐤p\in N_{\mathbf{k}} and αM\alpha\in M is locally nilpotent if and only if

  1. (i)(i)

    αi(I)\alpha\in\mathcal{R}_{i}(I) and p=λρip=\lambda\cdot\rho_{i} with λ𝐤\lambda\in\mathbf{k}; or

  2. (ii)(ii)

    αS\alpha\in S and (0α)supp(I)(\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I)\neq\emptyset; or

  3. (iii)(iii)

    αS\alpha\in S, (0α)supp(I)=(\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I)=\emptyset, and (m+0α)supp(I)(m+\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I)\neq\emptyset for all mSpm\in S\setminus p^{\bot}.

Proof.

The result follows directly from Lemmas Lemma 3.3, Lemma 3.4, and Lemma 3.5. ∎

Remark 3.8.

Let SS be a semigroup and let I𝐤[S]I\subset\mathbf{k}[S] be a monomial ideal. In the case where SS is not isomorphic to 0n\mathbb{Z}_{\geq 0}^{n}, by Theorem 2.2 there exist monomial ideals II such that not all derivations are liftable. Some of these non-liftable derivations may be locally nilpotent. Indeed, the non-liftable derivations constructed in the proofs of Lemma 2.3 and Lemma 2.8 are locally nilpotent.

Example 3.9.

With the notation in Example 1.1 for n=2n=2, we present here examples of all the degrees of homogeneous locally nilpotent derivations for different monomial ideals II. Red points indicate the degrees described in Theorem 3.7 (i)(i). Green points indicate the degrees described in Theorem 3.7 (ii)(ii). Finally, blue points indicate the degrees described in Theorem 3.7 (iii)(iii).

Figure 7.

I1=(x2y5,x3y2,x5y)I_{1}=(x^{2}y^{5},x^{3}y^{2},x^{5}y)

Figure 8.

I2=(xy5,x3y2,x5)I_{2}=(xy^{5},x^{3}y^{2},x^{5})

Figure 9.

I3=(y5,x3y2,x5)I_{3}=(y^{5},x^{3}y^{2},x^{5})

Some derivations appearing in Theorem 3.7 are zero derivations. In the following lemma, we provide a criterion to determine when a homogeneous derivation of 𝐤[S]\mathbf{k}[S] is the trivial derivation of 𝐤[S]/I\mathbf{k}[S]/I.

Lemma 3.10.

A homogeneous derivation ¯α,p:𝐤[S]/I𝐤[S]/I\overline{\partial}_{\alpha,p}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I with pN𝐤p\in N_{\mathbf{k}} and αM\alpha\in M is trivial if and only if

((Sp)+α)supp(I).\big{(}(S\setminus p^{\bot})+\alpha\big{)}\subset\operatorname{supp}(I)\,.
Proof.

The derivation =α,p\partial=\partial_{\alpha,p} maps to the trivial derivation via π\pi if and only if for every mSm\in S we have (𝐱m)I\partial(\mathbf{x}^{m})\in I. This holds if and only if for every mSm\in S with p(m)0p(m)\neq 0, we have m+αm+\alpha contained in supp(I)\operatorname{supp}(I), so 𝐱m+αI\mathbf{x}^{m+\alpha}\in I. This last statement is equivalent to the one given in the lemma. ∎

Remark 3.11.

It is sufficient to verify the condition in Lemma 3.10 on a set of generators of SS.

4. Automorphism group of monomial algebras

In this section, we apply our classification of homogeneous locally nilpotent derivations on 𝐤[S]/I\mathbf{k}[S]/I to describe Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) in the case where II is a monomial ideal with cofinite support. Our approach is inspired by Demazure’s description of Aut(X)\operatorname{Aut}(X) for a complete smooth toric variety XX [Dem70], which was recently revisited in modern terms by one of the authors [LLA22]. In the following proposition, we prove that the automorphism group Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is linear algebraic.

Proposition 4.1.

Let SS be a semigroup and let II be a monomial ideal with cofinite support. Then the automorphism group of 𝐤[S]/I\mathbf{k}[S]/I is linear algebraic.

Proof.

Since the monomial algebra 𝐤[S]/I\mathbf{k}[S]/I is finite-dimensional as a vector space over 𝐤\mathbf{k} and every automorphism φ:𝐤[S]/I𝐤[S]/I\varphi\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I is also an invertible linear map, we have that Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) embeds in GL(𝐤[S]/I)\operatorname{GL}(\mathbf{k}[S]/I). Moreover, φGL(𝐤[S]/I)\varphi\in\operatorname{GL}(\mathbf{k}[S]/I) belongs to Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) if and only if φ(𝐱¯m+m)=φ(𝐱¯m)φ(𝐱¯m)\varphi(\overline{\mathbf{x}}^{m+m^{\prime}})=\varphi(\overline{\mathbf{x}}^{m})\cdot\varphi(\overline{\mathbf{x}}^{m^{\prime}}) for all m,mSm,m^{\prime}\in S, which is a closed condition. We conclude that Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is a closed subgroup of GL(𝐤[S]/I)\operatorname{GL}(\mathbf{k}[S]/I) and thus is linear algebraic. ∎

Let now T=Spec𝐤[M]T=\operatorname{Spec}\mathbf{k}[M]. Then, TT acts faithfully on 𝐤[S]\mathbf{k}[S] via the usual toric action given by the comorphism

𝐤[S]𝐤[M]𝐤[S]defined by𝐱m𝐱m𝐱m.\mathbf{k}[S]\to\mathbf{k}[M]\otimes\mathbf{k}[S]\quad\text{defined by}\quad\mathbf{x}^{m}\mapsto\mathbf{x}^{m}\otimes\mathbf{x}^{m}.

A closed point tTt\in T corresponds to a maximal ideal of 𝐤[M]\mathbf{k}[M] and, in turn, corresponds to a homomorphism t:M𝐤t\colon M\rightarrow\mathbf{k}^{*}. Indeed, tt induces a homomorphism 𝐤[M]𝐤\mathbf{k}[M]\rightarrow\mathbf{k} via 𝐱mt(m)\mathbf{x}^{m}\mapsto t(m), and the maximal ideal defining the point tt is the kernel of this homomorphism. With this notation, the action of a closed point tTt\in T on 𝐤[S]\mathbf{k}[S] is given by

𝐤[S]𝐤[S]defined by𝐱mt𝐱m=t(m)𝐱m.\displaystyle\mathbf{k}[S]\to\mathbf{k}[S]\quad\text{defined by}\quad\mathbf{x}^{m}\mapsto t\cdot\mathbf{x}^{m}=t(m)\mathbf{x}^{m}. (9)

In particular, the TT-action preserves every monomial ideal II and thus defines a TT-action on 𝐤[S]/I\mathbf{k}[S]/I.

Definition 4.2.

Let SS be a semigroup, and let II be a monomial ideal in 𝐤[S]\mathbf{k}[S]. We say that II is full if no ray generator of SS is contained in supp(I)\operatorname{supp}(I).

In the case that a monomial ideal II in 𝐤[S]\mathbf{k}[S] is not full, let SS^{\prime} be the smallest subsemigroup of SS containing Ssupp(I)S\setminus\operatorname{supp}(I), and let I=I𝐤[S]I^{\prime}=I\cap\mathbf{k}[S^{\prime}]. Then, we have that 𝐤[S]/I=𝐤[S]/I\mathbf{k}[S]/I=\mathbf{k}[S^{\prime}]/I^{\prime} and II^{\prime} is full.

Example 4.3.

If SS is the first octant as in Example 1.1, then the action of TT is given by component-wise multiplication. Furthermore, a monomial ideal I𝐤[S]I\subset\mathbf{k}[S] is full if and only if no variable xix_{i} belongs to II.

Lemma 4.4.

Let SS be a semigroup and let II be a monomial ideal. If II is full, then TT acts faithfully on 𝐤[S]/I\mathbf{k}[S]/I.

Proof.

Let tTt\in T act as the identity on 𝐤[S]/I\mathbf{k}[S]/I. By (9), we have t.𝐱m=t(m)𝐱m=𝐱mt.\mathbf{x}^{m}=t(m)\mathbf{x}^{m}=\mathbf{x}^{m} for all mSsupp(I)m\in S\setminus\operatorname{supp}(I). Thus, t(m)=1t(m)=1 for all mSsupp(I)m\in S\setminus\operatorname{supp}(I). If II is full, then Ssupp(I)S\setminus\operatorname{supp}(I) spans MM_{\mathbb{R}}. Therefore, t(m)=1t(m)=1 for all mMm\in M, and hence tt is the identity in TT. ∎

Under the condition of Lemma 4.4, we denote the image of TT inside Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) also by TT. In the next lemma, we show that, in certain cases, TT is a maximal torus in Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I).

Lemma 4.5.

Letting SS be the first octant, we let II be a full monomial ideal with cofinite support. We also let T=Spec𝐤[M]Aut𝐤(𝐤[S]/I)T=\operatorname{Spec}\mathbf{k}[M]\subset\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I). Then, the centralizer of TT equals TT. In particular, TT is a maximal torus of Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I).

Proof.

Let gAut𝐤(𝐤[S]/I)g\in\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) be defined by

g(𝐱¯m)=mSsupp(I)am,m𝐱¯mfor allmSsupp(I).g(\overline{\mathbf{x}}^{m^{\prime}})=\sum_{m\in S\setminus\operatorname{supp}(I)}a_{m^{\prime},m}\cdot\overline{\mathbf{x}}^{m}\quad\mbox{for all}\quad m^{\prime}\in S\setminus\operatorname{supp}(I)\,.

Assume that gg belongs to the centralizer of TT, that is, tg(𝐱¯m)=gt(𝐱¯m)t\circ g(\overline{\mathbf{x}}^{m^{\prime}})=g\circ t(\overline{\mathbf{x}}^{m^{\prime}}), for all tTt\in T. This implies

mSsupp(I)t(m)am,m𝐱¯m=mSsupp(I)t(m)am,m𝐱¯m,for all tT.\sum_{m\in S\setminus\operatorname{supp}(I)}t(m)\cdot a_{m^{\prime},m}\cdot\overline{\mathbf{x}}^{m}=\sum_{m\in S\setminus\operatorname{supp}(I)}t(m^{\prime})\cdot a_{m^{\prime},m}\cdot\overline{\mathbf{x}}^{m},\quad\mbox{for all }t\in T\,.

This equality can only hold if am,m=0a_{m^{\prime},m}=0 for all mmm\neq m^{\prime}. Letting am,m=ama_{m,m}=a_{m}, we conclude

g(𝐱¯m)=am𝐱¯mfor allmSsupp(I).\displaystyle g(\overline{\mathbf{x}}^{m})=a_{m}\overline{\mathbf{x}}^{m}\quad\mbox{for all}\quad m\in S\setminus\operatorname{supp}(I)\,. (10)

Finally, since the basis vector eie_{i} belongs to Ssupp(I)S\setminus\operatorname{supp}(I) as SS is the first octant and II is full, for every m=(m1,,mn)Mm=(m_{1},\ldots,m_{n})\in M, we have that am=aeimia_{m}=\prod a_{e_{i}}^{m_{i}} and so gTg\in T by (9). This proves, in particular, that the centralizer of TT is TT itself, thus TT is a maximal torus of Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I). ∎

The hypothesis that SS is the first octant in the above lemma is essential, as the following example shows.

Example 4.6.

Let SS be the semigroup 0{(1,0),(1,1),(1,2)}\mathbb{Z}_{\geq 0}\{(1,0),(1,1),(1,2)\}. We consider the monomial ideal I𝐤[S]I\subset\mathbf{k}[S] given by I=(𝐱(2,0),𝐱(2,1),𝐱(2,2),𝐱(2,3),𝐱(2,4))I=(\mathbf{x}^{(2,0)},\mathbf{x}^{(2,1)},\mathbf{x}^{(2,2)},\mathbf{x}^{(2,3)},\mathbf{x}^{(2,4)}). Thus, 𝐤[S]/I𝐤𝐱¯(0,0)𝐤𝐱¯(1,0)𝐤𝐱¯(1,1)𝐤𝐱¯(1,2)\mathbf{k}[S]/I\simeq\mathbf{k}\cdot\overline{\mathbf{x}}^{(0,0)}\oplus\mathbf{k}\cdot\overline{\mathbf{x}}^{(1,0)}\oplus\mathbf{k}\cdot\overline{\mathbf{x}}^{(1,1)}\oplus\mathbf{k}\cdot\overline{\mathbf{x}}^{(1,2)}. If gg is an element of the centralizer of (𝐤)2(\mathbf{k}^{*})^{2}, then, using the same argument as in the proof of Lemma 4.5 up to (10), we conclude that

g(𝐱¯(1,0))=a(1,0)𝐱¯(1,0),g(𝐱¯(1,1))=a(1,1)𝐱¯(1,1),andg(𝐱¯(1,2))=a(1,2)𝐱¯(1,2).g(\overline{\mathbf{x}}^{(1,0)})=a_{(1,0)}\overline{\mathbf{x}}^{(1,0)},\quad g(\overline{\mathbf{x}}^{(1,1)})=a_{(1,1)}\overline{\mathbf{x}}^{(1,1)},\quad\text{and}\quad g(\overline{\mathbf{x}}^{(1,2)})=a_{(1,2)}\overline{\mathbf{x}}^{(1,2)}.

However, these three coefficients a(1,0)a_{(1,0)}, a(1,1)a_{(1,1)}, and a(1,2)a_{(1,2)} are algebraically independent since the relation a(1,0)a(1,2)=a(1,1)2a_{(1,0)}\cdot a_{(1,2)}=a^{2}_{(1,1)} does not hold in the quotient. This implies that there is a 3-dimensional torus acting faithfully on 𝐤[S]/I\mathbf{k}[S]/I that is an extension of TT. In particular, the centralizer of TT is larger than TT.

Furthermore, a straightforward verification shows that 𝐤[S]/I\mathbf{k}[S]/I is isomorphic to 𝐤[S]/I\mathbf{k}[S^{\prime}]/I^{\prime} with SS^{\prime} being the first octant in M=3M^{\prime}=\mathbb{Z}^{3} and I=(𝐱(2,0,0),𝐱(0,2,0),𝐱(0,0,2),𝐱(1,1,0),𝐱(1,0,1),𝐱(0,1,1))I^{\prime}=(\mathbf{x}^{(2,0,0)},\mathbf{x}^{(0,2,0)},\mathbf{x}^{(0,0,2)},\mathbf{x}^{(1,1,0)},\mathbf{x}^{(1,0,1)},\mathbf{x}^{(0,1,1)}). Now, Lemma 4.5 shows that a maximal torus of Aut𝐤(𝐤[S]/I)Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S^{\prime}]/I^{\prime})\simeq\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is indeed 3-dimensional.

Definition 4.7 (See [RvS21, Definition 7.1]).

Let GG be a linear algebraic group, and let TGT\subset G be a maximal torus. An algebraic subgroup UGU\subset G of dimension mm, isomorphic to (𝔾a)m(\mathbb{G}_{a})^{m}, is called a generalized root subgroup with respect to TT if there exists a character 𝐱α\mathbf{x}^{\alpha} of TT, referred to as the weight of UU, such that

tε(s)t1=ε(𝐱α(t)s)for all tT and all s(𝔾a)m,t\circ\varepsilon(s)\circ t^{-1}=\varepsilon(\mathbf{x}^{\alpha}(t)\cdot s)\quad\mbox{for all }t\in T\mbox{ and all }s\in(\mathbb{G}_{a})^{m}\,,

where ε:(𝔾a)mU\varepsilon\colon(\mathbb{G}_{a})^{m}\to U is a fixed isomorphism. The weight of a generalized root subgroup UU does not depend on the choice of the isomorphism ε\varepsilon. A generalized root subgroup is said to be maximal if it is maximal by inclusion among generalized root subgroups.

With this definition, we arrive at the following well-known result. In lack of a reference, we provide a proof.

Proposition 4.8.

Let GG be a connected linear algebraic group, and let TT be a maximal torus of GG. Then GG is generated as a group by TT and all maximal generalized root subgroups.

Proof.

Since the characteristic of 𝐤\mathbf{k} is zero, there exists a Levi subgroup LGL\subset G, i.e., a reductive subgroup such that G=LRG=L\ltimes R, where RR is the unipotent radical of GG [Bor91, Definition 11.22]. Furthermore, we can assume that TLT\subset L. Moreover, all generalized root subgroups of LL with respect to TT are one-dimensional, and together with TT, they generate LL [Spr09, Proposition 8.1.1.]. Now, let gRg\in R. Since G=LRG=L\ltimes R and TLT\subset L, we have that RR is a direct product of root subgroups R1,,RkR_{1},\dots,R_{k} with respect to TT with some weight 𝐱αi\mathbf{x}^{\alpha_{i}}, i=1,,ki=1,\dots,k. Thus, each RiR_{i} belongs to the maximal generalized root subgroup corresponding to 𝐱αi\mathbf{x}^{\alpha_{i}}. This completes the proof of the proposition. ∎

In the next lemma, we show that elements in generalized root subgroups correspond to homogeneous locally nilpotent derivations.

Lemma 4.9.

Let SS be the first octant, and let II be a full monomial ideal with cofinite support. Suppose gAut𝐤0(𝐤[S]/I)g\in\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I) and that gg is contained in a generalized root subgroup with weight 𝐱α\mathbf{x}^{\alpha} with respect to TT. Then g=exp¯g=\exp\overline{\partial}, where ¯\overline{\partial} is a locally nilpotent derivation of degree α\alpha.

Proof.

Since gg is contained in a generalized root subgroup of Aut𝐤0(𝐤[S]/I)\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I), gg is unipotent, so the smallest algebraic subgroup UU containing gg is a root subgroup isomorphic to 𝔾a\mathbb{G}_{\mathrm{a}} with weight 𝐱α\mathbf{x}^{\alpha}. The injection UAut𝐤0(𝐤[S]/I)U\hookrightarrow\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I) induces a 𝔾a\mathbb{G}_{\mathrm{a}}-action on 𝐤[S]/I\mathbf{k}[S]/I. By (1), we have g=exp(s¯)g=\exp(s\overline{\partial}) for some locally nilpotent derivation ¯:𝐤[S]/I𝐤[S]/I\overline{\partial}\colon\mathbf{k}[S]/I\to\mathbf{k}[S]/I and some s𝔾as\in\mathbb{G}_{\mathrm{a}}. Since s¯s\overline{\partial} is again locally nilpotent, we can assume g=exp¯g=\exp\overline{\partial} by replacing ¯\overline{\partial} with s¯s\overline{\partial}. According to Definition 4.7,

texp(s¯)t1=exp(𝐱α(t)s¯)for alltT and all s𝔾a.t\circ\exp(s\overline{\partial})\circ t^{-1}=\exp(\mathbf{x}^{\alpha}(t)s\overline{\partial})\quad\text{for all}\quad t\in T\text{ and all }s\in\mathbb{G}_{a}.

Since conjugation commutes with the exponential, we obtain

t¯t1=𝐱α(t)¯for alltT.\displaystyle t\circ\overline{\partial}\circ t^{-1}=\mathbf{x}^{\alpha}(t)\cdot\overline{\partial}\quad\text{for all}\quad t\in T. (11)

Now, letting mSsupp(I)m^{\prime}\in S\setminus\operatorname{supp}(I), we write

(𝐱¯m)=mSsupp(I)am𝐱¯m.\partial(\overline{\mathbf{x}}^{m^{\prime}})=\sum_{m\in S\setminus\operatorname{supp}(I)}a_{m}\overline{\mathbf{x}}^{m}.

Applying (11) to 𝐱¯m\overline{\mathbf{x}}^{m^{\prime}}, we get

t1(m)mSsupp(I)t(m)am𝐱¯m=t(α)mSsupp(I)am𝐱¯m.t^{-1}(m^{\prime})\sum_{m\in S\setminus\operatorname{supp}(I)}t(m)\cdot a_{m}\overline{\mathbf{x}}^{m}=t(\alpha)\sum_{m\in S\setminus\operatorname{supp}(I)}a_{m}\overline{\mathbf{x}}^{m}.

This equality can only hold if am=0a_{m}=0 for all but possibly one value of mm. This implies that (𝐱¯m)=am𝐱¯m\partial(\overline{\mathbf{x}}^{m^{\prime}})=a_{m}\overline{\mathbf{x}}^{m}, and thus ¯\overline{\partial} is homogeneous. ∎

Remark 4.10.

Non-trivial locally nilpotent derivations on 𝐤[S]/I\mathbf{k}[S]/I were classified in Theorem 3.7 and Lemma 3.10. Note that in the case where II has cofinite support, only conditions (i)(i) and (ii)(ii) can hold in Theorem 3.7 since (0α)supp(I)(\mathbb{Z}_{\geq 0}\cdot\alpha)\cap\operatorname{supp}(I) is never empty for αS\alpha\in S.

Let (S,I)\mathcal{R}(S,I) denote the subset of MM consisting of α\alpha for which there exists a non-zero homogeneous locally nilpotent derivation ¯α,p\overline{\partial}_{\alpha,p}. According to Lemma 4.9, these are precisely the weights of the non-trivial generalized root subgroups of Aut𝐤0(𝐤[S]/I)\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I). Moreover, for each α(S,I)\alpha\in\mathcal{R}(S,I), we define

G(α)={pN𝐤¯α,p is locally nilpotent}.G(\alpha)=\{p\in N_{\mathbf{k}}\mid\overline{\partial}_{\alpha,p}\text{ is locally nilpotent}\}.

Next, we define

ια:G(α)Aut𝐤(𝐤[S]/I)bypexp¯α,p.\displaystyle\iota_{\alpha}\colon G(\alpha)\to\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)\quad\text{by}\quad p\mapsto\exp\overline{\partial}_{\alpha,p}\,. (12)

As shown in Lemma 4.9, the image of each ια\iota_{\alpha} is the maximal generalized root subgroup associated with the weight 𝐱α\mathbf{x}^{\alpha}. Since exp¯=id\exp\overline{\partial}=\operatorname{id} if and only if ¯=0\overline{\partial}=0, it follows that ια\iota_{\alpha} is injective, implying that the maximal generalized root subgroup corresponding to the weight 𝐱α\mathbf{x}^{\alpha} is isomorphic to the group structure of the vector space G(α)G(\alpha).

Now, we can present our main result regarding the connected component of Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I), which directly follows from Lemma 4.9 and the aforementioned considerations.

Proposition 4.11.

Let SS be the first octant and II be a full monomial ideal with cofinite support. Also, let T=Spec𝐤[M]Aut𝐤(𝐤[S]/I)T=\operatorname{Spec}\mathbf{k}[M]\subset\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I). Then, the neutral component Aut𝐤0(𝐤[S]/I)\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I) is spanned by TT and the images of ια\iota_{\alpha} for all α(S,I)\alpha\in\mathcal{R}(S,I).

Remark 4.12.

When XX is a complete toric variety, all the generalized root subgroups of Aut0(X)\operatorname{Aut}^{0}(X) are 1-dimensional [Dem70] (see also [LLA22, Theorem 2 (i)(i)]). However, this property does not hold for the automorphism group of monomial algebras, as illustrated by the following example.

Example 4.13.

Let M=2M=\mathbb{Z}^{2} and S=02S=\mathbb{Z}^{2}_{\geq 0}. Consider the ideal I=(y2,x2y,x3)I=(y^{2},x^{2}y,x^{3}). By Theorem 3.7 and Lemma 3.10, we have that α=(1,0)(S,I)\alpha=(1,0)\in\mathcal{R}(S,I) and G(α)=N𝐤G(\alpha)=N_{\mathbf{k}}. Furthermore, every p=(p1,p2)G(α)p=(p_{1},p_{2})\in G(\alpha) defines the following derivation and automorphisms, respectively.

¯p,α1:𝐤[S]/I𝐤[S]/Ix¯p1x¯2y¯p2x¯y¯so thatexp(¯p,α1):𝐤[S]/I𝐤[S]/Ix¯x¯+p1x¯2y¯y¯+p2x¯y¯.\begin{array}[]{ccc}\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{1}}\colon\mathbf{k}[S]/I&\to&\mathbf{k}[S]/I\\ \overline{x}&\mapsto&p_{1}\overline{x}^{2}\\ \overline{y}&\mapsto&p_{2}\overline{x}\overline{y}\\ \end{array}&\quad\mbox{so that}&\begin{array}[]{rcl}\operatorname{exp}(\overline{\partial}_{p,\alpha_{1}})\colon\mathbf{k}[S]/I&\to&\mathbf{k}[S]/I\\ \overline{x}&\mapsto&\overline{x}+p_{1}\overline{x}^{2}\\ \overline{y}&\mapsto&\overline{y}+p_{2}\overline{x}\overline{y}\\ \end{array}\\ \end{array}\,.

Since ια\iota_{\alpha} is injective, the maximal generalized root subgroup with weight 𝐱α\mathbf{x}^{\alpha} is 2-dimensional.

Definition 4.14.

Let SS and SS^{\prime} be semigroups and let ISI\subset S and ISI^{\prime}\subset S^{\prime} be monomial ideals. A 𝐤\mathbf{k}-algebra homomorphism g:𝐤[S]/I𝐤[S]/Ig\colon\mathbf{k}[S]/I\to\mathbf{k}[S^{\prime}]/I^{\prime} is called a toric morphism if there exists a semigroup homomorphism g^:SS\widehat{g}\colon S\to S^{\prime} with g^(supp(I))supp(I)\widehat{g}(\operatorname{supp}(I))\subset\operatorname{supp}(I^{\prime}) such that g(𝐱¯m)=𝐱¯g^(m)g(\overline{\mathbf{x}}^{m})=\overline{\mathbf{x}}^{\widehat{g}(m)}. A toric automorphism of 𝐤[S]/I\mathbf{k}[S]/I is a toric endomorphism that admits an inverse which is also toric. We denote the set of toric automorphisms of 𝐤[S]/I\mathbf{k}[S]/I by Aut(S,I)\operatorname{Aut}(S,I). An automorphism gAut(S,I)g\in\operatorname{Aut}(S,I) corresponds to an automorphism g^:SS\widehat{g}\colon S\to S such that g^(supp(I))=supp(I)\widehat{g}(\operatorname{supp}(I))=\operatorname{supp}(I).

Lemma 4.15.

The group Aut(S,I)\operatorname{Aut}(S,I) is finite.

Proof.

Let SS be a semigroup. Consider gAut(S,I)g\in\operatorname{Aut}(S,I), which corresponds to the semigroup automorphism g^:SS\widehat{g}\colon S\to S. The map g^\widehat{g} induces a permutation of the set of rays S(1)S(1). Consequently, we obtain a homomorphism from Aut(S,I)\operatorname{Aut}(S,I) to the symmetric group on S(1)S(1). Furthermore, since S(1)S(1) spans MM_{\mathbb{R}} as a vector space, the action of g^\widehat{g} on S(1)S(1) uniquely determines g^\widehat{g}, implying that the homomorphism is injective. ∎

Proposition 4.16.

If II is a full monomial ideal with cofinite support, then Aut𝐤(𝐤[S]/I)/Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)/\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) is a finite group generated by the images of toric morphisms.

Proof.

Since Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is a linear algebraic group by Proposition 4.1, it follows that the quotient group F=Aut𝐤(𝐤[S]/I)/Aut𝐤0(𝐤[S]/I)F=\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)/\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) is finite. Let gAut𝐤(𝐤[S]/I)g\in\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) and let [g][g] be its image in FF. As all maximal tori are conjugate in a linear algebraic group, there exists hAut𝐤0(𝐤[S]/I)h\in\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) such that

(gh)1T(gh)=h1(g1Tg)h=T.(g\circ h)^{-1}\circ T\circ(g\circ h)=h^{-1}\circ(g^{-1}\circ T\circ g)\circ h=T\,.

Since [g]=[gh][g]=[g\circ h], we can replace gg with ghg\circ h, assuming that gg normalizes TT.

Now, for mSsupp(I)m^{\prime}\in S\setminus\operatorname{supp}(I), we have

g(𝐱¯m)=mSsupp(I)am𝐱¯m.g(\overline{\mathbf{x}}^{m^{\prime}})=\sum_{m\in S\setminus\operatorname{supp}(I)}a_{m}\overline{\mathbf{x}}^{m}\,.

The set V=𝐤𝐱¯m𝐤[S]/IV=\mathbf{k}\cdot\overline{\mathbf{x}}^{m^{\prime}}\subset\mathbf{k}[S]/I is a 1-dimensional vector subspace of 𝐤[S]/I\mathbf{k}[S]/I. Applying gT=Tgg\circ T=T\circ g to VV, we obtain

𝐤mSsupp(I)am𝐱¯m=𝐤{mSsupp(I)t(m)am𝐱¯mtT}.\displaystyle\mathbf{k}\cdot\sum_{m\in S\setminus\operatorname{supp}(I)}a_{m}\overline{\mathbf{x}}^{m}=\mathbf{k}\cdot\left\{\sum_{m\in S\setminus\operatorname{supp}(I)}t(m)a_{m}\overline{\mathbf{x}}^{m}\mid t\in T\right\}\,.

The right-hand side is a 𝐤\mathbf{k}-vector space, while the left-hand side is a 𝐤\mathbf{k}-vector space if and only if am=0a_{m}=0 for all but at most one mSsupp(I)m\in S\setminus\operatorname{supp}(I). This implies that g(𝐱¯m)=am𝐱¯mg(\overline{\mathbf{x}}^{m^{\prime}})=a_{m}\overline{\mathbf{x}}^{m} for some mSsupp(I)m\in S\setminus\operatorname{supp}(I). Replacing gg by gtg\circ t with an appropriate choice of tTt\in T, we obtain [g]=[gt][g]=[g\circ t] and g(𝐱¯m)=𝐱¯mg(\overline{\mathbf{x}}^{m^{\prime}})=\overline{\mathbf{x}}^{m}. Finally, since II is full, the basis vectors eiS=0ne_{i}\in S=\mathbb{Z}_{\geq 0}^{n} are not in the support of II. Hence, we can define a unique homomorphism g^:SS\widehat{g}\colon S\to S given by g^(ei)=mi\widehat{g}(e_{i})=m_{i} where g(𝐱¯ei)=𝐱¯mig(\overline{\mathbf{x}}^{e_{i}})=\overline{\mathbf{x}}^{m_{i}}. Since gg is an automorphism, g^\widehat{g} is also an automorphism, and thus gg is a toric automorphism. ∎

We can now present our main theorem, which summarizes all the results in this section and provides a description of the automorphism group of certain monomial algebras. Recall that the map ια\iota_{\alpha} is defined in (12).

Theorem 4.17.

Let SS be the first octant, and let II be a full monomial ideal with cofinite support. Also, let T=Spec𝐤[M]Aut𝐤(𝐤[S]/I)T=\operatorname{Spec}\mathbf{k}[M]\subset\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I).

  1. (i)(i)

    The automorphism group Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is linear algebraic, and TT is a maximal torus of Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I).

  2. (ii)(ii)

    The neutral component Aut𝐤0(𝐤[S]/I)\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I) is generated by TT and the image of ια\iota_{\alpha} for all α(S,I)\alpha\in\mathcal{R}(S,I).

  3. (iii)(iii)

    The quotient group Aut𝐤(𝐤[S]/I)/Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)/\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I) is generated by the images of toric automorphisms.

Proof.

Statement (i)(i) is contained in Proposition 4.1 and Lemma 4.5. Statement (ii)(ii) is in Proposition 4.11. Finally, statement (iii)(iii) is in Proposition 4.16. ∎

We finish the paper providing several examples illustrating our results.

Example 4.18.

Letting M=M=\mathbb{Z} and S=0S=\mathbb{Z}_{\geq 0}, we let I=(x4)I=(x^{4}). The group of toric automorphisms is trivial so

Aut𝐤(𝐤[S]/I)=Aut𝐤0(𝐤[S]/I)=Aut𝐤(𝐤[x]/(x4)).\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)=\operatorname{Aut}^{0}_{\mathbf{k}}(\mathbf{k}[S]/I)=\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[x]/(x^{4}))\,.

By Theorem 3.7 and Lemma 3.10, we have that only α1=1\alpha_{1}=1 and α2=2\alpha_{2}=2 give rise to non-trivial homogeneous locally nilpotent derivations different from the zero derivation on 𝐤[S]/I\mathbf{k}[S]/I. For pN𝐤=G(αi)=𝐤p\in N_{\mathbf{k}}=G(\alpha_{i})=\mathbf{k}, for i=1,2i=1,2, we obtain the derivations

¯p,α1:𝐤[S]/I𝐤[S]/Ix¯px¯2and¯p,α2:𝐤[S]/I𝐤[S]/Ix¯px¯3.\begin{array}[]{ccc}\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{1}}\colon\mathbf{k}[S]/I&\to&\mathbf{k}[S]/I\\ \overline{x}&\mapsto&p\overline{x}^{2}\\ \end{array}&\quad\mbox{and}&\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{2}}\colon\mathbf{k}[S]/I&\to&\mathbf{k}[S]/I\\ \overline{x}&\mapsto&p\overline{x}^{3}\\ \end{array}\\ \end{array}\,.

Applying the exponential map, we obtain

exp(¯p,α1):𝐤[S]/I𝐤[S]/Ix¯x¯+px¯2+p2x¯3andexp(¯p,α2):𝐤[S]/I𝐤[S]/Ix¯x¯+px¯3.\begin{array}[]{ccc}\begin{array}[]{rcl}\exp(\overline{\partial}_{p,\alpha_{1}})\colon\mathbf{k}[S]/I&\to&\mathbf{k}[S]/I\\ \overline{x}&\mapsto&\overline{x}+p\overline{x}^{2}+p^{2}\overline{x}^{3}\\ \end{array}&\quad\mbox{and}&\begin{array}[]{rcl}\exp(\overline{\partial}_{p,\alpha_{2}})\colon\mathbf{k}[S]/I&\to&\mathbf{k}[S]/I\\ \overline{x}&\mapsto&\overline{x}+p\overline{x}^{3}\\ \end{array}\\ \end{array}\,.

Letting e1=1,e2=x¯,e3=x¯2,e4=x¯3e_{1}=1,e_{2}=\overline{x},e_{3}=\overline{x}^{2},e_{4}=\overline{x}^{3} we have that 𝐤[S]/I=𝐤e1𝐤e2𝐤e3𝐤e4\mathbf{k}[S]/I=\mathbf{k}\cdot e_{1}\oplus\mathbf{k}\cdot e_{2}\oplus\mathbf{k}\cdot e_{3}\oplus\mathbf{k}\cdot e_{4}. Moreover, Aut𝐤(𝐤[S]/I)GL4(𝐤)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)\subset\operatorname{GL}_{4}(\mathbf{k}) by the proof of Proposition 4.1. And by Theorem 4.17, the automorphism group Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is generated by the matrices

(10000t0000t20000t3),(100001000r100r22r1),and(1000010000100s01),\begin{array}[]{cccc}\begin{pmatrix}1&0&0&0\\ 0&t&0&0\\ 0&0&t^{2}&0\\ 0&0&0&t^{3}\\ \end{pmatrix},&\begin{pmatrix}1&0&0&0\\ 0&1&0&0\\ 0&r&1&0\\ 0&r^{2}&2r&1\\ \end{pmatrix},&\mbox{and}&\begin{pmatrix}1&0&0&0\\ 0&1&0&0\\ 0&0&1&0\\ 0&s&0&1\\ \end{pmatrix},\end{array}

with t𝔾mt\in\mathbb{G}_{\mathrm{m}}, r𝔾ar\in\mathbb{G}_{\mathrm{a}} and s𝔾as\in\mathbb{G}_{\mathrm{a}}.

Example 4.19.

Letting M=2M=\mathbb{Z}^{2} and S=02S=\mathbb{Z}_{\geq 0}^{2} we let I=(x2,y2)I=(x^{2},y^{2}). The group of toric automorphisms is F/2F\simeq\mathbb{Z}/2\mathbb{Z} with generator given by the permutation x¯y¯\overline{x}\mapsto\overline{y} and y¯x¯\overline{y}\to\overline{x}. Hence, Aut𝐤(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I) is generated by Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) and FF.

Figure 10. I=(x2,y2)I=(x^{2},y^{2})

By Theorem 3.7 and Lemma 3.10, we have that only α1=(1,0)\alpha_{1}=(1,0) and α2=(0,1)\alpha_{2}=(0,1) give rise to non-trivial homogeneous locally nilpotent derivations with G(α1)=𝐤(0,1)G(\alpha_{1})=\mathbf{k}\cdot(0,1) and G(α2)=𝐤(1,0)G(\alpha_{2})=\mathbf{k}\cdot(1,0). Letting p=(p1,p2)N𝐤p=(p_{1},p_{2})\in N_{\mathbf{k}}, this yields

¯p,α1:𝐤[x,y]/I𝐤[x,y]/Ix¯0y¯p2x¯y¯and¯p,α2:𝐤[x,y]/I𝐤[x,y]/Ix¯p1x¯y¯y¯0.\begin{array}[]{cc}\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{1}}\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&0\\ \overline{y}&\mapsto&p_{2}\overline{x}\overline{y}\\ \end{array}\par\quad\mbox{and}&\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{2}}\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&p_{1}\overline{x}\overline{y}\\ \overline{y}&\mapsto&0\\ \end{array}\\ \end{array}\,.

Applying the exponential map, we obtain

exp(¯p,α1):𝐤[x,y]/I𝐤[x,y]/Ix¯x¯y¯y¯+p2x¯y¯andexp(¯p,α2):𝐤[x,y]/I𝐤[x,y]/Ix¯x¯+p1x¯y¯y¯y¯.\begin{array}[]{cc}\begin{array}[]{rcl}\exp(\overline{\partial}_{p,\alpha_{1}})\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&\overline{x}\\ \overline{y}&\mapsto&\overline{y}+p_{2}\overline{x}\overline{y}\\ \end{array}\par\quad\mbox{and}&\begin{array}[]{rcl}\exp(\overline{\partial}_{p,\alpha_{2}})\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&\overline{x}+p_{1}\overline{x}\overline{y}\\ \overline{y}&\mapsto&\overline{y}\\ \end{array}\\ \end{array}\,.

Letting e1=1,e2=y¯,e3=x¯,e4=x¯y¯e_{1}=1,e_{2}=\overline{y},e_{3}=\overline{x},e_{4}=\overline{x}\overline{y} we have that 𝐤[S]/I=𝐤e1𝐤e2𝐤e3𝐤e4\mathbf{k}[S]/I=\mathbf{k}\cdot e_{1}\oplus\mathbf{k}\cdot e_{2}\oplus\mathbf{k}\cdot e_{3}\oplus\mathbf{k}\cdot e_{4}. Moreover Aut𝐤(𝐤[x,y]/I)GL4(𝐤)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[x,y]/I)\subset\operatorname{GL}_{4}(\mathbf{k}) by the proof of Proposition 4.1. And by Theorem 4.17, Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) is generated by the matrices

(10000t20000t10000t1t2),(1000010000100r01),and(10000100001000s1),\begin{array}[]{ccc}\begin{pmatrix}1&0&0&0\\ 0&t_{2}&0&0\\ 0&0&t_{1}&0\\ 0&0&0&t_{1}t_{2}\\ \end{pmatrix},&\begin{pmatrix}1&0&0&0\\ 0&1&0&0\\ 0&0&1&0\\ 0&r&0&1\\ \end{pmatrix},\quad\mbox{and}&\begin{pmatrix}1&0&0&0\\ 0&1&0&0\\ 0&0&1&0\\ 0&0&s&1\\ \end{pmatrix},\end{array}

with (t1,t2)T=𝔾m2(t_{1},t_{2})\in T=\mathbb{G}_{\mathrm{m}}^{2}, r𝔾ar\in\mathbb{G}_{\mathrm{a}} and s𝔾as\in\mathbb{G}_{\mathrm{a}}.

Finally, all matrices are lower triangular, so the generator of FF is not contained in Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I). We conclude that

Aut𝐤(𝐤[S]/I)Aut𝐤0(𝐤[S]/I)F.\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)\simeq\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I)\ltimes F\,.
Example 4.20.

Letting M=2M=\mathbb{Z}^{2} and S=02S=\mathbb{Z}_{\geq 0}^{2} we let I=(x2,xy,y2)I=(x^{2},xy,y^{2}).


Figure 11. I=(x2,xy,y2)I=(x^{2},xy,y^{2})

The group of toric automorphisms FF is again isomorphic to /2\mathbb{Z}/2\mathbb{Z} with generator given by the permutation x¯y¯\overline{x}\mapsto\overline{y} and y¯x¯\overline{y}\mapsto\overline{x}. In this case, as we will show by the end of this example that FAut𝐤0(𝐤[S]/I)F\subset\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I). Hence,

Aut𝐤(𝐤[S]/I)=Aut𝐤0(𝐤[S]/I).\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[S]/I)=\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I).

By Theorem 3.7 and Lemma 3.10, we have that only α1=(1,1)\alpha_{1}=(1,-1) and α2=(1,1)\alpha_{2}=(-1,1) give rise to non-trivial homogeneous locally nilpotent derivations with G(α1)=𝐤(0,1)G(\alpha_{1})=\mathbf{k}\cdot(0,1) and G(α2)=𝐤(1,0)G(\alpha_{2})=\mathbf{k}\cdot(1,0). Letting p=(p1,p2)N𝐤p=(p_{1},p_{2})\in N_{\mathbf{k}}, this yield

¯p,α1:𝐤[x,y]/I𝐤[x,y]/Ix¯0y¯p2x¯and¯p,α2:𝐤[x,y]/I𝐤[x,y]/Ix¯p1y¯y¯0.\begin{array}[]{cc}\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{1}}\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&0\\ \overline{y}&\mapsto&p_{2}\overline{x}\\ \end{array}\par\quad\mbox{and}&\begin{array}[]{rcl}\overline{\partial}_{p,\alpha_{2}}\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&p_{1}\overline{y}\\ \overline{y}&\mapsto&0\\ \end{array}\\ \end{array}\,.

Applying the exponential map, we obtain:

exp(¯p,α1):𝐤[x,y]/I𝐤[x,y]/Ix¯x¯y¯y¯+p2x¯andexp(¯p,α2):𝐤[x,y]/I𝐤[x,y]/Ix¯x¯+p1y¯y¯y¯.\begin{array}[]{cc}\begin{array}[]{rcl}\exp(\overline{\partial}_{p,\alpha_{1}})\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&\overline{x}\\ \overline{y}&\mapsto&\overline{y}+p_{2}\overline{x}\\ \end{array}\par\quad\mbox{and}&\begin{array}[]{rcl}\exp(\overline{\partial}_{p,\alpha_{2}})\colon\mathbf{k}[x,y]/I&\to&\mathbf{k}[x,y]/I\\ \overline{x}&\mapsto&\overline{x}+p_{1}\overline{y}\\ \overline{y}&\mapsto&\overline{y}\\ \end{array}\\ \end{array}\,.

Letting now e1=1,e2=y¯e_{1}=1,e_{2}=\overline{y} and e3=x¯e_{3}=\overline{x} we have that 𝐤[S]/I=𝐤e1𝐤e2𝐤e3\mathbf{k}[S]/I=\mathbf{k}\cdot e_{1}\oplus\mathbf{k}\cdot e_{2}\oplus\mathbf{k}\cdot e_{3}. Moreover Aut𝐤(𝐤[x,y]/I)GL3(𝐤)\operatorname{Aut}_{\mathbf{k}}(\mathbf{k}[x,y]/I)\subset\operatorname{GL}_{3}(\mathbf{k}) by the proof of Proposition 4.1. And by Theorem 4.17, Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) is generated by the matrices

(1000t2000t1),(1000100r1),and(10001s001),\begin{array}[]{ccc}\begin{pmatrix}1&0&0\\ 0&t_{2}&0\\ 0&0&t_{1}\\ \end{pmatrix},&\begin{pmatrix}1&0&0\\ 0&1&0\\ 0&r&1\\ \end{pmatrix},\quad\mbox{and}&\begin{pmatrix}1&0&0\\ 0&1&s\\ 0&0&1\\ \end{pmatrix},\end{array}

with (t1,t2)T=𝔾m2(t_{1},t_{2})\in T=\mathbb{G}_{\mathrm{m}}^{2}, r𝔾ar\in\mathbb{G}_{\mathrm{a}} and s𝔾as\in\mathbb{G}_{\mathrm{a}}. The group Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) is isomorphic to GL2(𝐤)\operatorname{GL}_{2}(\mathbf{k}) taking the lower right 2×22\times 2 blocks of the above matrices, c.f. [MS05, Corollary 2.2].

In this case, FAut𝐤0(𝐤[S]/I)F\subset\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) since the permutation x¯y¯\overline{x}\mapsto\overline{y}, y¯x¯\overline{y}\mapsto\overline{x} is obtained as φ1ψφ\varphi^{-1}\circ\psi\circ\varphi where φ,ψAut𝐤0(𝐤[S]/I)\varphi,\psi\in\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) are defined via φ(x¯,y¯)=(x¯,x¯y¯)\varphi(\overline{x},\overline{y})=(\overline{x},\overline{x}-\overline{y}) and ψ(x¯,y¯)=(x¯y¯,y¯)\psi(\overline{x},\overline{y})=(\overline{x}-\overline{y},\overline{y}). This phenomenon where some toric morphisms are contained in Aut𝐤0(𝐤[S]/I)\operatorname{Aut}_{\mathbf{k}}^{0}(\mathbf{k}[S]/I) can only arise when there exist non-trivial root subgroups with weights α\alpha and α-\alpha.

References

  • [Bar02] Alexander Barvinok. A course in convexity, volume 54 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2002.
  • [Bor91] Armand Borel. Linear algebraic groups, volume 126 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1991.
  • [BS95] Paulo Brumatti and Aron Simis. The module of derivations of a Stanley-Reisner ring. Proc. Amer. Math. Soc., 123:1309–1318, 1995.
  • [CLS11] David A. Cox, John B. Little, and Henry K. Schenck. Toric varieties, volume 124. American Mathematical Soc., 2011.
  • [Dai03] Daniel Daigle. Locally nilpotent derivations, lecture notes for the september school of algebraic geometry, 2003.
  • [DCEP82] Corrado De Concini, David Eisenbud, and Claudio Procesi. Hodge algebras, volume 91 of Astérisque. Société Mathématique de France, Paris, 1982. With a French summary.
  • [Dem70] Michel Demazure. Sous-groupes algébriques de rang maximum du groupe de cremona. Annales scientifiques de l’École normale supérieure, 3(4):507–588, 1970.
  • [DL24] Roberto Díaz and Alvaro Liendo. On the Automorphism Group of Non-Necessarily Normal Affine Toric Varieties. Int. Math. Res. Not. IMRN, 2:1624–1649, 2024.
  • [Fre17] Gene Freudenburg. Algebraic theory of locally nilpotent derivations, volume 136 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, second edition, 2017. Invariant Theory and Algebraic Transformation Groups, VII.
  • [Ful93] William Fulton. Introduction to toric varieties. Number 131 in Annals of mathematics studies. Princeton university press, 1993.
  • [HH11] Jürgen Herzog and Takayuki Hibi. Monomial ideals, volume 260 of Graduate Texts in Mathematics. Springer-Verlag London, Ltd., London, 2011.
  • [KLL15] Frank Kutzschebauch, Matthias Leuenberger, and Alvaro Liendo. The algebraic density property for affine toric varieties. J. Pure Appl. Algebra, 219(8):3685–3700, 2015.
  • [Kra17] Hanspeter Kraft. On the lie algebra of vector fields of affine varieties. https://dmi.unibas.ch/fileadmin/user_upload/dmi/Personen/Kraft_Hanspeter/Lie_Algebra_Vector_Fields.pdf, 2017.
  • [Lie10] Alvaro Liendo. Affine 𝕋\mathbb{T}-varieties of complexity one and locally nilpotent derivations. Transformation Groups, 15(2):389–425, 2010.
  • [LLA22] Alvaro Liendo and Giancarlo Lucchini Arteche. Automorphisms of products of toric varieties. Math. Res. Lett., 29(2):529–540, 2022.
  • [MS05] Ezra Miller and Bernd Sturmfels. Combinatorial commutative algebra, volume 227 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2005.
  • [Oda83] Tadao Oda. Convex bodies and algebraic geometry: an introduction to the theory of toric varieties. Springer, 1983.
  • [PR23] Alexander Perepechko and Andriy Regeta. When is the automorphism group of an afine variety nested? Transf. Groups, 28:401–412, 2023.
  • [PR24] Alexander Perepechko and Andriy Regeta. Automorphism groups of affine varieties consisting of algebraic elements. Proc. Amer. Math. Soc, 152:2377–2383, 2024.
  • [Rei76] Gerald Allen Reisner. Cohen-Macaulay quotients of polynomial rings. Advances in Math., 21(1):30–49, 1976.
  • [RvS21] Andriy Regeta and Immanuel van Santen. Characterizing smooth affine spherical varieties via the automorphism group. Journal de l’École polytechnique—Mathématiques, 8:379–414, 2021.
  • [Spr09] T. A. Springer. Linear algebraic groups. Modern Birkhäuser Classics. Birkhäuser Boston, Inc., Boston, MA, second edition, 2009.
  • [Sta75] Richard P. Stanley. The upper bound conjecture and Cohen-Macaulay rings. Studies in Appl. Math., 54(2):135–142, 1975.
  • [Tad09] Yohannes Tadesse. Derivations preserving a monomial ideal. Proc. Amer. Math. Soc., 137(9):2935–2942, 2009.