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

Waring and cactus ranks and Strong Lefschetz Property for annihilators of symmetric forms

Mats Boij Department of Mathematics, KTH Royal Institute of Technology, S-100 44 Stockholm, Sweden [email protected] Juan Migliore Department of Mathematics, University of Notre Dame, Notre Dame, IN 46556, USA [email protected] Rosa M. Miró-Roig Facultat de Matemàtiques i Informàtica. Universitat de Barcelona. Gran Via 585. 08007 Barcelona. Spain [email protected]  and  Uwe Nagel Department of Mathematics, University of Kentucky, 715 Patterson Office Tower, Lexington, KY 40506-0027, USA [email protected]
Abstract.

In this note we show that the complete symmetric polynomials are dual generators of compressed artinian Gorenstein algebras satisfying the Strong Lefschetz Property. This is the first example of an explicit dual form with these properties.

For complete symmetric forms of any degree in any number of variables, we provide an upper bound for the Waring rank by establishing an explicit power sum decomposition.

Moreover, we determine the Waring rank, the cactus rank, the resolution and the Strong Lefschetz Property for any Gorenstein algebra defined by a symmetric cubic form. In particular, we show that the difference between the Waring rank and the cactus rank of a symmetric cubic form can be made arbitrarily large by increasing the number of variables.

We provide upper bounds for the Waring rank of generic symmetric forms of degrees four and five.

The authors want to thank CIRM in Trento, MFO in Oberwolfach and the University of Kentucky for support and hospitality. The second author was partially supported by a Simons Foundation grant #309556. The third author was partially supported by MTM2016-78623-P. The fourth author was partially supported by Simons Foundation grants #317096 and #636513.

1. Introduction

Let kk be a field of characteristic 0 and consider the polynomial ring S=k[x1,x2,,xn]S=k[x_{1},x_{2},\dots,x_{n}] and its inverse system E=k[X1,X2,,Xn]E=k[X_{1},X_{2},\dots,X_{n}]; more precisely, EE is an SS-module where xix_{i} acts as Xi\frac{\partial}{\partial X_{i}}. It is a well-known fact ([IK, Lemma 3.12]) that if ff is a general polynomial of degree dd in EE then the annihilator J=ann(f)J=\hbox{ann}(f) in SS defines an artinian Gorenstein algebra S/JS/J with compressed Hilbert function, meaning that the “first half” of the Hilbert function of S/JS/J agrees with the Hilbert function of SS, and the second half is of course determined by symmetry.

This begs the question of how “general” ff has to be in order to obtain a compressed Hilbert function in this way; in particular, can one exhibit a specific polynomial and prove that it has this property? This problem arose in conversation during a Research in Pairs involving the authors, in Trento in 2009, while working on a different project. At that time, computer experiments showed that the complete symmetric polynomial (i.e. the sum of all the monomials of degree dd) has this property. In the intervening years the authors were able to prove this fact, and in the process many related and intriguing problems presented themselves. This paper is the result.

In Section 2, in Theorem 2.11 we give a proof of the fact just mentioned, that when f=hdf=h_{d} is the complete symmetric polynomial then S/ann(hd)S/\hbox{ann}(h_{d}) has compressed Hilbert function. (The fact that S/ann(f)S/\hbox{ann}(f) has compressed Hilbert function was also proven by F. Gesmundo and J. M. Landsberg [GL], with a different method.) In fact, Theorem 2.11 proves not only this, but that S/ann(hd)S/\hbox{ann}(h_{d}) even satisfies the Strong Lefschetz Property (SLP). We recall that a graded artinian kk-algebra AA satisfies the Weak Lefschetz Property (WLP) if multiplication by a general linear form ×\times\ell from any component to the next has maximal rank, and it satisfies SLP if maximal rank also holds for ×j:[A]i[A]i+j\times\ell^{j}:[A]_{i}\rightarrow[A]_{i+j} for all i0i\geq 0 and j0j\geq 0. In the same section we also give a useful decomposition of ann(hd)\hbox{ann}(h_{d}) (Theorem 2.12), separating into the cases where dd is even or odd. (The result for even degree is a bit stronger.)

One of the main topics of this paper concerns the Waring rank, and we begin looking at this in Section 3. If ff is a homogeneous polynomial of degree mm, recall that its Waring rank is the smallest kk so that ff can be written in the form L1m++LkmL_{1}^{m}+\dots+L_{k}^{m} for linear forms L1,,LkL_{1},\dots,L_{k}, and a power sum decomposition is any expression of ff in this form (not necessarily minimal). Let us denote by hn,mh_{n,m} the complete symmetric polynomial of degree mm in nn variables. The main result of Section 3 is a specific power sum decomposition for hn,2d+1h_{n,2d+1} and separately for hn,2dh_{n,2d}, for any nn and dd. In both the even and odd cases, our power sum decomposition has (n+dd)\binom{n+d}{d} terms, giving an upper bound for the Waring rank. Thus our bound is a polynomial of degree dd in nn. Note that the Waring rank for general forms of degree δ\delta grows like a polynomial of degree δ1\delta-1 in nn (see Remark 3.3). Later in the paper we will show that our bound for hn,3h_{n,3} is sharp (see Theorem 3.2), and we believe that it is sharp for all odd degrees. In the case of even degrees, we note that fewer terms may be possible at least in some cases (see Remark 3.3).

In Section 4, we consider only symmetric cubic polynomials. This is, in a sense, the heart of this paper. By choosing a very special basis for the vector space of symmetric forms of degree 3, we can view the family of all such polynomials as being parameterized by a projective plane, 2\mathbb{P}^{2}. Inside of this plane, we consider the forms that have Waring rank at most nn. We show that this locus can be viewed as a rational cuspidal cubic curve, 𝒞\mathcal{C}, and we give its equation (Lemma 4.2). From our perspective there are two important points on 𝒞\mathcal{C}: the (unique) flex point 𝒬\mathcal{Q} and the cusp 𝒫\mathcal{P}. Then there are three important lines: the line 1\ell_{1} joining 𝒫\mathcal{P} and 𝒬\mathcal{Q}, the tangent line 2\ell_{2} to 𝒞\mathcal{C} at 𝒫\mathcal{P} (meaning the unique line that meets 𝒞\mathcal{C} at 𝒫\mathcal{P} and no other point), and the tangent 3\ell_{3} to 𝒞\mathcal{C} at 𝒬\mathcal{Q}. We give the equations of these three lines.

With this preparation, we divide the points of 2\mathbb{P}^{2} as follows: 𝒫\mathcal{P}, 𝒬\mathcal{Q}, the remaining points of 𝒞\mathcal{C}, the remaining points of 1\ell_{1}, the remaining points of 2\ell_{2}, the remaining points of 3\ell_{3}, and the remaining points of 2\mathbb{P}^{2}. For each point of 2\mathbb{P}^{2}, depending on which of these categories it falls into, we give the Hilbert function of the corresponding algebra, a description of the generators of ann(F)\hbox{ann}(F), the minimal free resolution, and both the Waring rank and the cactus rank of the defining symmetric polynomial. Recall that the cactus rank of a form FF is the smallest degree of a zero-dimensional subscheme whose saturated ideal is contained in ann(F)\hbox{ann}(F). We show as a consequence that for symmetric cubic forms, the Waring rank and the cactus rank are equal except for points on 2\ell_{2} other than 𝒫\mathcal{P} (Corollary 4.9). We end this section by showing that for any symmetric cubic form FF, the corresponding algebra S/ann(F)S/\hbox{ann}(F) satisfies the SLP (Proposition 4.10).

In the last section, we look at the symmetric generic Waring rank, that is the Waring rank of a generic symmetric form. We provide upper bounds for this generic rank for quartics and quintics, generalizing the results obtained for cubics in the previous section. These bounds are close to the lower bounds given by the Hilbert function, but we cannot establish the actual symmetric generic rank.

2. The SLP for the artinian Gorenstein algebra defined by a complete symmetric polynomial

Let kk be a field of characteristic 0 and consider the polynomial ring S=k[x1,x2,,xn]S=k[x_{1},x_{2},\dots,x_{n}] and its inverse system E=k[X1,X2,,Xn]E=k[X_{1},X_{2},\dots,X_{n}] which means that EE is an SS-module where xix_{i} acts as Xi\frac{\partial}{\partial X_{i}}.

The homogeneous, or complete, symmetric polynomials are defined as

hd(X1,X2,,Xn)=i¯n,|i¯|=dX1i1X2i2Xnin.h_{d}(X_{1},X_{2},\dots,X_{n})=\sum_{\underline{i}\in\mathbb{N}^{n},|\underline{i}|=d}X_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}}.

The goal of this section is to prove that the Gorenstein algebra S/ann(he)S/\operatorname{ann}(h_{e}) is compressed and satisfies the SLP (Theorem 2.11). The fact that it is compressed was also shown in [GL].

We fix nn and in the formal power series ring k[[X1,X2,,Xn]]k[[X_{1},X_{2},\dots,X_{n}]], which is also a graded SS-module, we define H=d=0hd=i¯nX1i1X2i2XninH=\sum_{d=0}^{\infty}h_{d}=\sum_{\underline{i}\in\mathbb{N}^{n}}X_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}}. We will often use \ell to denote the linear form x1+x2++xnx_{1}+x_{2}+\dots+x_{n} in SS.

Lemma 2.1.

The inverse system of the ideal ()=(x1+x2++xn)(\ell)=(x_{1}+x_{2}+\dots+x_{n}) is given by the subring E=k[X1Xn,X2Xn,,Xn1Xn]E^{\prime}=k[X_{1}-X_{n},X_{2}-X_{n},\dots,X_{n-1}-X_{n}] considered as an SS-module.

Proof.

The linear form \ell annihilates any polynomial in X1Xn,X2Xn,,Xn1XnX_{1}-X_{n},X_{2}-X_{n},\dots,X_{n-1}-X_{n} and the dimension of the inverse system is (n2+dn2)\binom{n-2+d}{n-2} in degree dd which equals the dimension of EdE^{\prime}_{d}. Thus EE^{\prime} is the inverse system of the ideal ()(\ell). ∎

Lemma 2.2.

hd=(n+d1)hd1.\ell\circ h_{d}=(n+d-1)h_{d-1}.

Proof.

The monomial X1i1X2i2XninX_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}} of degree d1d-1 occurs from the derivatives xjXjX1i1X2i2Xnin=(ij+1)X1i1X2i2Xninx_{j}\circ X_{j}X_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}}=(i_{j}+1)X_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}}. The sum of these contributions is n+ij=n+d1n+\sum i_{j}=n+d-1. ∎

Definition 2.3.

Let Φ:SE\Phi\colon S\longrightarrow E be the kk-linear map defined on the monomial basis by

Φ(x1i1x2i2xnin)=i1!i2!in!X1i1X2i2Xnin.\Phi(x_{1}^{i_{1}}x_{2}^{i_{2}}\cdots x_{n}^{i_{n}})=i_{1}!i_{2}!\cdots i_{n}!X_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}}.

for each i¯=(i1,i2,,in)n\underline{i}=(i_{1},i_{2},\dots,i_{n})\in\mathbb{N}^{n}.

Remark 2.4.

Note that the inverse image of the monomial basis of EdE_{d} gives the dual basis with respect to the action. Moreover, powers of \ell are sent to multiples of the homogeneous symmetric polynomials, Φ(d)=d!hd\Phi(\ell^{d})=d!h_{d}.

Definition 2.5.

For any a¯=(a1,a2,,an1)n1\underline{a}=(a_{1},a_{2},\dots,a_{n-1})\in\mathbb{N}^{n-1} we define

Fa¯=(X1Xn)a1(X2Xn)a2(Xn1Xn)an1andfa¯=Φ1(Fa¯).F_{\underline{a}}=(X_{1}-X_{n})^{a_{1}}(X_{2}-X_{n})^{a_{2}}\cdots(X_{n-1}-X_{n})^{a_{n-1}}\quad\text{and}\quad f_{\underline{a}}=\Phi^{-1}(F_{\underline{a}}).
Lemma 2.6.

For any a¯=(a1,a2,,an1)\underline{a}=(a_{1},a_{2},\dots,a_{n-1}) we have that

fa¯H=Fa¯(1X1)1+a1(1X2)1+a2(1Xn1)1+an1(1Xn)1+i=1n1ai.f_{\underline{a}}\circ H=\frac{F_{\underline{a}}}{(1-X_{1})^{1+a_{1}}(1-X_{2})^{1+a_{2}}\cdots(1-X_{n-1})^{1+a_{n-1}}(1-X_{n})^{1+\sum_{i=1}^{n-1}a_{i}}}.
Proof.

We expand fa¯f_{\underline{a}} by the binomial theorem as

Φ1(j¯n1,j¯a¯(1)i=1n1aiji(a1j1)(a2j2)(an1jn1)X1j1X2j2Xn1jn1Xni=1n1aiji)=j¯n1,j¯a¯(1)i=1n1aiji(a1j1)(a2j2)(an1jn1)x1j1x2j2xn1jn1xni=1n1aijij1!j2!jn1!(i=1n1aiji)!.\Phi^{-1}\left(\sum_{\underline{j}\in\mathbb{N}^{n-1},\underline{j}\leq\underline{a}}(-1)^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\binom{a_{1}}{j_{1}}\binom{a_{2}}{j_{2}}\cdots\binom{a_{n-1}}{j_{n-1}}X_{1}^{j_{1}}X_{2}^{j_{2}}\cdots X_{n-1}^{j_{n-1}}X_{n}^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\right)\\ =\sum_{\underline{j}\in\mathbb{N}^{n-1},\underline{j}\leq\underline{a}}(-1)^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\binom{a_{1}}{j_{1}}\binom{a_{2}}{j_{2}}\cdots\binom{a_{n-1}}{j_{n-1}}\frac{x_{1}^{j_{1}}x_{2}^{j_{2}}\cdots x_{n-1}^{j_{n-1}}x_{n}^{\sum_{i=1}^{n-1}a_{i}-j_{i}}}{j_{1}!j_{2}!\cdots j_{n-1}!(\sum_{i=1}^{n-1}a_{i}-j_{i})!}.\\

Now look at the contribution to fa¯Hf_{\underline{a}}\circ H by each term of this sum. The only monomials in HH that contribute are the multiples of the corresponding monomial in the variables X1,X2,,XnX_{1},X_{2},\dots,X_{n}. We get that

x1j1x2j2xn1jn1xni=1n1aijiH=x1j1x2j2xn1jn1xni=1n1aijiX1j1X2j2Xn1jn1Xni=1n1aijiH=b¯nx1j1x2j2xn1jn1xni=1n1aijiX1b1+j1X2b2+j2Xn1bn1+jn1Xnbn+i=1n1aiji=b¯n(b1+j1)!b1!(b2+j2)!b2!(bn1+jn1)!bn1!(bn+aiji)!bn!X1b1X2b2Xnbn=j1!(1X1)1+j1j2!(1X2)1+j2jn1!(1Xn1)1+jn1(aiji)!)(1Xn)1+i=1n1aiji.x_{1}^{j_{1}}x_{2}^{j_{2}}\cdots x_{n-1}^{j_{n-1}}x_{n}^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\circ H=x_{1}^{j_{1}}x_{2}^{j_{2}}\cdots x_{n-1}^{j_{n-1}}x_{n}^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\circ X_{1}^{j_{1}}X_{2}^{j_{2}}\cdots X_{n-1}^{j_{n-1}}X_{n}^{\sum_{i=1}^{n-1}a_{i}-j_{i}}H\\ =\sum_{\underline{b}\in\mathbb{N}^{n}}x_{1}^{j_{1}}x_{2}^{j_{2}}\cdots x_{n-1}^{j_{n-1}}x_{n}^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\circ X_{1}^{b_{1}+j_{1}}X_{2}^{b_{2}+j_{2}}\cdots X_{n-1}^{b_{n-1}+j_{n-1}}X_{n}^{b_{n}+\sum_{i=1}^{n-1}a_{i}-j_{i}}\\ =\sum_{\underline{b}\in\mathbb{N}^{n}}\frac{(b_{1}+j_{1})!}{b_{1}!}\frac{(b_{2}+j_{2})!}{b_{2}!}\cdots\frac{(b_{n-1}+j_{n-1})!}{b_{n-1}!}\frac{(b_{n}+\sum a_{i}-j_{i})!}{b_{n}!}X_{1}^{b_{1}}X_{2}^{b_{2}}\cdots X_{n}^{b_{n}}\\ =\frac{j_{1}!}{(1-X_{1})^{1+j_{1}}}\frac{j_{2}!}{(1-X_{2})^{1+j_{2}}}\cdots\frac{j_{n-1}!}{(1-X_{n-1})^{1+j_{n-1}}}\frac{(\sum a_{i}-j_{i})!)}{(1-X_{n})^{1+\sum_{i=1}^{n-1}a_{i}-j_{i}}}.

Summing over the terms of fa¯f_{\underline{a}} now gives

fa¯H=j¯n1,j¯a¯(1)i=1n1aiji(a1j1)(a2j2)(an1jn1)(1X1)1+j1(1X2)1+j2(1Xn1)1+jn1(1Xn)1+aiji=(11Xn1X1)a1(11Xn1X2)a2(11Xn1Xn1)an11(1Xn)ai(1)i=1n1ai(1Xi)=(1)i=1n1ai(XnX1)a1(1X1)1+a1(XnX2)a2(1X2)1+a2(XnXn1)an1(1Xn1)1+an11(1Xn)1+ai=Fa¯(1X1)1+a1(1X2)1+a2(1Xn1)1+an1(1Xn)1+i=1n1aif_{\underline{a}}\circ H=\sum_{\underline{j}\in\mathbb{N}^{n-1},\underline{j}\leq\underline{a}}\frac{(-1)^{\sum_{i=1}^{n-1}a_{i}-j_{i}}\binom{a_{1}}{j_{1}}\binom{a_{2}}{j_{2}}\cdots\binom{a_{n-1}}{j_{n-1}}}{(1-X_{1})^{1+j_{1}}(1-X_{2})^{1+j_{2}}\cdots(1-X_{n-1})^{1+j_{n-1}}(1-X_{n})^{1+\sum a_{i}-j_{i}}}\\ =\left(1-\frac{1-X_{n}}{1-X_{1}}\right)^{a_{1}}\left(1-\frac{1-X_{n}}{1-X_{2}}\right)^{a_{2}}\cdots\left(1-\frac{1-X_{n}}{1-X_{n-1}}\right)^{a_{n-1}}\frac{1}{(1-X_{n})^{\sum a_{i}}}\frac{(-1)^{\sum_{i=1}^{n-1}a_{i}}}{\prod(1-X_{i})}\\ =(-1)^{\sum_{i=1}^{n-1}a_{i}}\frac{(X_{n}-X_{1})^{a_{1}}}{(1-X_{1})^{1+a_{1}}}\frac{(X_{n}-X_{2})^{a_{2}}}{(1-X_{2})^{1+a_{2}}}\cdots\frac{(X_{n}-X_{n-1})^{a_{n-1}}}{(1-X_{n-1})^{1+a_{n-1}}}\frac{1}{(1-X_{n})^{1+\sum a_{i}}}\\ =\frac{F_{\underline{a}}}{(1-X_{1})^{1+a_{1}}(1-X_{2})^{1+a_{2}}\cdots(1-X_{n-1})^{1+a_{n-1}}(1-X_{n})^{1+\sum_{i=1}^{n-1}a_{i}}}

which concludes the proof of the lemma. ∎

Lemma 2.7.

For a¯=(a1,a2,,an1)\underline{a}=(a_{1},a_{2},\dots,a_{n-1}) in n1\mathbb{N}^{n-1} with |a¯|=d|\underline{a}|=d we have that

  1. (1)

    fa¯hi=0f_{\underline{a}}\circ h_{i}=0, for i<2di<2d.

  2. (2)

    fa¯h2d=Fa¯f_{\underline{a}}\circ h_{2d}=F_{\underline{a}}.

  3. (3)

    fa¯h2d+m=Fa¯Gm,a¯f_{\underline{a}}\circ h_{2d+m}=F_{\underline{a}}G_{m,\underline{a}}, where

    Gm,a¯=i¯n,|i¯|=m(a1+i1i1)(a2+i2i2)(an1+in1in1)(aj+inin)X1i1X2i2Xnin.G_{m,\underline{a}}=\sum_{\underline{i}\in\mathbb{N}^{n},|\underline{i}|=m}\binom{a_{1}+i_{1}}{i_{1}}\binom{a_{2}+i_{2}}{i_{2}}\cdots\binom{a_{n-1}+i_{n-1}}{i_{n-1}}\binom{\sum a_{j}+i_{n}}{i_{n}}X_{1}^{i_{1}}X_{2}^{i_{2}}\cdots X_{n}^{i_{n}}.
Proof.

The first two statements follow from the fact that the initial term in fa¯h2df_{\underline{a}}\circ h_{2d} is Fa¯F_{\underline{a}} which has degree dd. The last statement is given by the power series expansion of

1(1X1)1+a1(1X2)1+a2(1Xn1)1+an1(1Xn)1+i=1n1ai.\frac{1}{(1-X_{1})^{1+a_{1}}(1-X_{2})^{1+a_{2}}\cdots(1-X_{n-1})^{1+a_{n-1}}(1-X_{n})^{1+\sum_{i=1}^{n-1}a_{i}}}.

Lemma 2.8.

For a¯=(a1,a2,,an1)\underline{a}=(a_{1},a_{2},\dots,a_{n-1}) and b¯=(b1,b2,,bn1)\underline{b}=(b_{1},b_{2},\dots,b_{n-1}) in n1\mathbb{N}^{n-1} with |a¯|=|b¯|=d|\underline{a}|=|\underline{b}|=d, we have that

fa¯Fb¯=i=1n1(ai+biai).f_{\underline{a}}\circ F_{\underline{b}}=\prod_{i=1}^{n-1}\binom{a_{i}+b_{i}}{a_{i}}.
Proof.

We have

fa¯Fb¯=i¯𝐍n1k=1n1xjikik!(akik)(xn)d|i¯|(d|i¯|)!j¯𝐍n1k=1n1xkjk(bkjk)(xn)d|j¯|=i¯𝐍n1k=1n1(akik)(bkik)=k=1n1i=0(aki)(bki)=i=1n1(ai+biai)f_{\underline{a}}\circ F_{\underline{b}}=\sum_{\underline{i}\in\mathbf{N}^{n-1}}\prod_{k=1}^{n-1}\frac{x_{j}^{i_{k}}}{i_{k}!}\binom{a_{k}}{i_{k}}\frac{(-x_{n})^{d-|\underline{i}|}}{(d-|\underline{i}|)!}\circ\sum_{\underline{j}\in\mathbf{N}^{n-1}}\prod_{k=1}^{n-1}x_{k}^{j_{k}}\binom{b_{k}}{j_{k}}(-x_{n})^{d-|\underline{j}|}\\ =\sum_{\underline{i}\in\mathbf{N}^{n-1}}\prod_{k=1}^{n-1}\binom{a_{k}}{i_{k}}\binom{b_{k}}{i_{k}}=\prod_{k=1}^{n-1}\sum_{i=0}^{\infty}\binom{a_{k}}{i}\binom{b_{k}}{i}=\prod_{i=1}^{n-1}\binom{a_{i}+b_{i}}{a_{i}}

where we use that

i=0(ai)(bi)=(a+ba)\sum_{i=0}^{\infty}\binom{a}{i}\binom{b}{i}=\binom{a+b}{a}

which comes from looking at the coefficient of tat^{a} in (1+t)a+b=(1+t)a(1+t)b(1+t)^{a+b}=(1+t)^{a}(1+t)^{b}. ∎

Definition 2.9.

For integers d0d\geq 0 let Md=Φ1(Ed)={fa¯:a¯n1,|a¯|=d}.M_{d}=\Phi^{-1}(E^{\prime}_{d})=\langle\{f_{\underline{a}}\colon\underline{a}\in\mathbb{N}^{n-1},|\underline{a}|=d\}\rangle.

Proposition 2.10.

The pairing iMdi×jMdjk\ell^{i}M_{d-i}\times\ell^{j}M_{d-j}\longrightarrow k given by f,g=(fg)h2d\langle f,g\rangle=(fg)\circ h_{2d} is trivial when iji\neq j and is perfect when i=ji=j.

Proof.

We have that a basis for iMdi\ell^{i}M_{d-i} is given by ifa¯\ell^{i}f_{\underline{a}} with |a¯|=di|\underline{a}|=d-i. We have that

(ifa¯jfb¯)h2d=(n+2d1)!(n+2dij1)!fa¯fb¯h2dij.(\ell^{i}f_{\underline{a}}\cdot\ell^{j}f_{\underline{b}})\circ h_{2d}=\frac{(n+2d-1)!}{(n+2d-i-j-1)!}f_{\underline{a}}f_{\underline{b}}\circ h_{2d-i-j}.

If iji\neq j, we have that either fa¯f_{\underline{a}} or fb¯f_{\underline{b}} annihilates h2dijh_{2d-i-j} by Lemma 2.7(1) since either 2(di)>2dij2(d-i)>2d-i-j or 2(dj)>2dij2(d-j)>2d-i-j. Hence the pairing is trivial.

If i=ji=j, it suffices to show that the pairing Mdi×MdikM_{d-i}\times M_{d-i}\longrightarrow k given by f,g=(fg)h2(di)\langle f,g\rangle=(fg)\circ h_{2(d-i)} is perfect. By Lemma 2.7 (2) we get that for ff in MdiM_{d-i} we have Φ(f)=fh2(di)\Phi(f)=f\circ h_{2(d-i)}. Hence the pairing is perfect since Φ:SdiEdi\Phi\colon S_{d-i}\longrightarrow E_{d-i} is invertible. ∎

Theorem 2.11.

For each e0e\geq 0 the Gorenstein algebra A=S/ann(he)A=S/\operatorname{ann}(h_{e}) is compressed and satisfies the SLP.

Proof.

For even socle degree, e=2de=2d, we need to show that the ideal ann(h2d)\operatorname{ann}(h_{2d}) is zero in degree dd in order to conclude that the algebra is compressed. In degree dd, we have that Md+Md1++dM0SdM_{d}+\ell M_{d-1}+\cdots+\ell^{d}M_{0}\subseteq S_{d}. Using Proposition 2.10 we see that the pairing given by f,g=(fg)h2d\langle f,g\rangle=(fg)\circ h_{2d} is perfect on each of the summands and trivial between any two of them. Since the sum of the dimensions of the summands equals the dimension of SdS_{d}, we conclude that the pairing is perfect on SdS_{d}, which implies that ann(h2d)\operatorname{ann}(h_{2d}) is trivial in degree dd.

For odd socle degree, e=2d+1e=2d+1, it is enough to show that ann(h2d+1)\operatorname{ann}(h_{2d+1}) is zero in degree dd. If ff has degree dd and satisfies fh2d+1=0f\circ h_{2d+1}=0 we also have (f)h2d+2(\ell f)\circ h_{2d+2} by Lemma 2.2, but we have shown that ann(h2d+2)\operatorname{ann}(h_{2d+2}) is trivial in degree d+1d+1 and therefore f=f=0\ell f=f=0. Hence the algebra is compressed.

Let A=S/ann(he)A=S/\operatorname{ann}(h_{e}). In order to prove that AA has the strong Lefschetz property, it is sufficient to show that e2i:AiAei\ell^{e-2i}\colon A_{i}\longrightarrow A_{e-i} is an isomorphism for all ie/2i\leq e/2. Assume that fSif\in S_{i} satisfies e2if=0\ell^{e-2i}f=0 in AeiA_{e-i}. This means that e2ifhe=0\ell^{e-2i}f\circ h_{e}=0, which by Lemma 2.2 means that fh2i=0f\circ h_{2i}=0. However, since A/ann(h2i)A/\operatorname{ann}(h_{2i}) is compressed, this means that f=0f=0 and the multiplication by e2i\ell^{e-2i} is injective and by symmetry of the Hilbert function also bijective. ∎

Theorem 2.12.

For even socle degree, e=2de=2d, we have that ann(h2d)=(Md+1Md)\operatorname{ann}(h_{2d})=(M_{d+1}\oplus\ell M_{d}) and for odd socle degree, e=2d+1e=2d+1, we have that ann(h2d+1)=(Md+1+2Md)\operatorname{ann}(h_{2d+1})=(M_{d+1}+\ell^{2}M_{d}).

Proof.

Since the algebra A=S/ann(h2d)A=S/\operatorname{ann}(h_{2d}) is compressed of even socle degree we have generators of ann(h2d)\operatorname{ann}(h_{2d}) only in degree d+1d+1 and the number of generators is given by

(n+dn1)(n+d2n1)=(n+d1n2)+(n+d2n2).\binom{n+d}{n-1}-\binom{n+d-2}{n-1}=\binom{n+d-1}{n-2}+\binom{n+d-2}{n-2}.

Both Md+1M_{d+1} and Md\ell M_{d} annihilate h2dh_{2d} and their intersection is trivial as we can see using the pairing given by f,g=(fg)h2d+2\langle f,g\rangle=(fg)\circ h_{2d+2}. Indeed, from Proposition 2.10 we have that for a non-zero element ff in Md+1M_{d+1}, there is an element gMd+1g\in M_{d+1} with f,g0\langle f,g\rangle\neq 0, while if fMdf\in\ell M_{d}, we have to have f,g=0\langle f,g\rangle=0 from the same proposition.

Moreover, dimkMd+1=(n+d1n2)\dim_{k}M_{d+1}=\binom{n+d-1}{n-2} and dimkMd=(n+d2n2)\dim_{k}\ell M_{d}=\binom{n+d-2}{n-2}, which proves that ann(h2d)=(Md+1Md)\operatorname{ann}(h_{2d})=(M_{d+1}\oplus\ell M_{d}).

Since the algebra A=S/ann(h2d+1)A=S/\operatorname{ann}(h_{2d+1}) is compressed of odd socle degree and we can have generators in degree d+1d+1 and d+2d+2. The dimension of ann(h2d+1)\operatorname{ann}(h_{2d+1}) in degree d+1d+1 is given by (n+dn1)(n+d1n1)=(n+d1n2)\binom{n+d}{n-1}-\binom{n+d-1}{n-1}=\binom{n+d-1}{n-2} which equals dimkMd+1\dim_{k}M_{d+1}. Since Md+1M_{d+1} annihilates h2d+1h_{2d+1} we must have that ann(h2d+1)\operatorname{ann}(h_{2d+1}) in degree d+1d+1 equals Md+1M_{d+1}. Let gSd+2g\in S_{d+2} be an element of ann(h2d+1)\operatorname{ann}(h_{2d+1}). Since we have that Md+1()d+1=Sd+1M_{d+1}\oplus(\ell)_{d+1}=S_{d+1} we must have that g=f1+f2g=\ell f_{1}+f_{2} for some f1Sd+1f_{1}\in S_{d+1} and f2f_{2} in (Md+1)d+2(M_{d+1})_{d+2}. Now gh2d+1=0g\circ h_{2d+1}=0 implies that f1h2d=0f_{1}\circ h_{2d}=0. From above it follows that ann(h2d)d+1=Md+1Md\operatorname{ann}(h_{2d})_{d+1}=M_{d+1}\oplus\ell M_{d} so f1Md+Md+1f_{1}\in\ell M_{d}+M_{d+1}. Thus we see that g2Md+(Md+1)g\in\ell^{2}M_{d}+(M_{d+1}), which concludes the proof. ∎

3. A power sum decomposition for the complete symmetric polynomial

In this section we will give a decomposition of the complete symmetric polynomial as a sum of powers of linear forms. The number of terms in this decomposition is in general much lower than the Waring rank of a general polynomial of the same degree (see Remark 3.3) and it seems likely that our decomposition is a Waring decomposition when the number of variables is higher than the degree and the degree is odd. We will prove this in the case of degree three.

Lemma 3.1.

For a polynomial f[X1,X2,,Xn]f\in\mathbb{Q}[X_{1},X_{2},\dots,X_{n}] we have that

f=0{f|Xn=0=0, and(x1+x2++xn)f=0.f=0\qquad\Longleftrightarrow\qquad\begin{cases}f|_{X_{n}=0}=0,\text{ and}\\ (x_{1}+x_{2}+\dots+x_{n})\circ f=0.\end{cases}
Proof.

Assume that f|Xn=0=0f|_{X_{n}=0}=0 and f=0\ell\circ f=0, where =x1+x2++xn\ell=x_{1}+x_{2}+\cdots+x_{n}. Then f=0f=0 or f=Xndgf=X_{n}^{d}g, where g|Xn=00g|_{X_{n}=0}\neq 0 and d>0d>0. Applying f=0\ell\circ f=0 we get

dXnd1g+Xndg=0dX_{n}^{d-1}g+X_{n}^{d}\ell\circ g=0

and division by Xnd1X_{n}^{d-1} shows that g|Xn=0=0g|_{X_{n}=0}=0. Hence f=0f=0. The other direction is trivial. ∎

In the following we will use the notation hn,dh_{n,d} for the complete symmetric polynomial of degree dd in the variables X1,X2,,XnX_{1},X_{2},\dots,X_{n}.

Theorem 3.2.

For n1n\geq 1 and d0d\geq 0 the following power sum decompositions hold for the complete symmetric functions

hn,2d+1=122d(2d+1)!k=0d(1)dk(n+2ddk)1i1i2ikn(hn,1+2j=1kXij)2d+1h_{n,2d+1}=\frac{1}{2^{2d}(2d+1)!}\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}\sum_{1\leq i_{1}\leq i_{2}\leq\cdots\leq i_{k}\leq n}\left(h_{n,1}+2\sum_{j=1}^{k}X_{i_{j}}\right)^{2d+1}

and

hn,2d=122d(2d)!k=0d(1)dk[(n+2d1dk)(n+2d1dk1)]1i1i2ikn(hn,1+2j=1kXij)2dh_{n,2d}=\frac{1}{2^{2d}(2d)!}\sum_{k=0}^{d}(-1)^{d-k}\left[\binom{n+2d-1}{d-k}-\binom{n+2d-1}{d-k-1}\right]\sum_{1\leq i_{1}\leq i_{2}\leq\cdots\leq i_{k}\leq n}\left(h_{n,1}+2\sum_{j=1}^{k}X_{i_{j}}\right)^{2d}

where in both cases the second summation contains just one term hn,12d+1h_{n,1}^{2d+1} and hn,12dh_{n,1}^{2d}, respectively, for k=0k=0.

Proof.

We will prove this by induction on nn and dd and we introduce the notation

Sn,k,d=1i1i2ikn(hn,1+2j=1kXij)dS_{n,k,d}=\sum_{1\leq i_{1}\leq i_{2}\leq\cdots\leq i_{k}\leq n}\left(h_{n,1}+2\sum_{j=1}^{k}X_{i_{j}}\right)^{d}

as an expression valid in any polynomial ring [X1,X2,,XN]\mathbb{Q}[X_{1},X_{2},\dots,X_{N}] with NnN\geq n. Furthermore, for m0m\geq 0, we introduce the notation

An,d,m=k=0d(1)dk(n+2ddk)Sn,k,2m+1Bn,d,m=k=0d(1)dk[(n+2d1dk)(n+2d1dk1)]Sn,k,2m\begin{split}A_{n,d,m}&=\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}S_{n,k,2m+1}\\ B_{n,d,m}&=\sum_{k=0}^{d}(-1)^{d-k}\left[\binom{n+2d-1}{d-k}-\binom{n+2d-1}{d-k-1}\right]S_{n,k,2m}\end{split}

for n1n\geq 1, d0d\geq 0 and m0m\geq 0. We shall prove by induction also on mdm\leq d that

(1) An,d,m={22d(2d+1)!hn,2d+1,m=d,0,m<dandBn,d,m={22d(2d)!hn,2d,m=d,0,m<d.A_{n,d,m}=\begin{cases}2^{2d}(2d+1)!h_{n,2d+1},&m=d,\\ 0,&m<d\end{cases}\quad\text{and}\quad B_{n,d,m}=\begin{cases}2^{2d}(2d)!h_{n,2d},&m=d,\\ 0,&m<d.\end{cases}

The statement of the theorem is (1) in the case m=dm=d.

The base of the induction is given by the cases d=0d=0, n=1n=1 and m=0m=0. For d=0d=0 we have that m=0m=0 and we get

An,0,0=Sn,0,1=i=1nXi=201!hn,1andBn,0,0=Sn,0,0=1=200!hn,0A_{n,0,0}=S_{n,0,1}=\sum_{i=1}^{n}X_{i}=2^{0}1!h_{n,1}\quad\text{and}\quad B_{n,0,0}=S_{n,0,0}=1=2^{0}0!h_{n,0}

so the assertions hold for d=0d=0. For n=1n=1 we have that S1,k,d=(1+2k)dX1dS_{1,k,d}=(1+2k)^{d}X_{1}^{d} and we get

A1,d,m=k=0d(1)dk(1+2ddk)(1+2k)2m+1X12m+1=k=0d(1)k(1+2dk)(1+2(dk))2m+1X12m+1=12k=02d+1(1)k(1+2dk)(1+2d2k)2m+1X12m+1={22d(2d+1)!X12d+1,m=d,0,0m<d\begin{split}A_{1,d,m}&=\sum_{k=0}^{d}(-1)^{d-k}\binom{1+2d}{d-k}(1+2k)^{2m+1}X_{1}^{2m+1}=\sum_{k=0}^{d}(-1)^{k}\binom{1+2d}{k}(1+2(d-k))^{2m+1}X_{1}^{2m+1}\\ &=\frac{1}{2}\sum_{k=0}^{2d+1}(-1)^{k}\binom{1+2d}{k}(1+2d-2k)^{2m+1}X_{1}^{2m+1}=\begin{cases}2^{2d}(2d+1)!X_{1}^{2d+1},&m=d,\\ 0,&0\leq m<d\end{cases}\end{split}

and

B1,d,m=1(2d+1)(2m+1)x1A1,d,m={22d(2d)!X12d,m=d,0,0m<dB_{1,d,m}=\frac{1}{(2d+1)(2m+1)}x_{1}\circ A_{1,d,m}=\begin{cases}2^{2d}(2d)!X_{1}^{2d},&m=d,\\ 0,&0\leq m<d\end{cases}

according to the previous equality. Thus both equalities hold for n=1n=1.

For m=0m=0 we have that

An,d,0=k=0d(1)dk(n+2ddk)Sn,k,1=hn,1k=0d(1)dk(n+2ddk)(n1+kk)n+2kn=hn,1k=0d(1)dk(n+2ddk)[(n+kk)+(n+k1k1)]\begin{split}A_{n,d,0}&=\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}S_{n,k,1}=h_{n,1}\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}\binom{n-1+k}{k}\frac{n+2k}{n}\\ &=h_{n,1}\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}\left[\binom{n+k}{k}+\binom{n+k-1}{k-1}\right]\end{split}

and the coefficient of hn,1h_{n,1} equals the coefficient of tdt^{d} in the formal power series

(1t)n+2d[1(1t)n+1+t(1t)n+1]=(1t)2d1+t(1t)2d1(1-t)^{n+2d}\left[\frac{1}{(1-t)^{n+1}}+\frac{t}{(1-t)^{n+1}}\right]=(1-t)^{2d-1}+t(1-t)^{2d-1}

which is equal to

(1)d[(2d1d)(2d1d1)]={1,d=0,0,d>0.(-1)^{d}\left[\binom{2d-1}{d}-\binom{2d-1}{d-1}\right]=\begin{cases}1,&d=0,\\ 0,&d>0.\end{cases}

For Bn,d,0B_{n,d,0} we get

Bn,d,0=k=0d(1)dk[(n+2d1dk)(n+2d1d1k)](n1+kk)B_{n,d,0}=\sum_{k=0}^{d}(-1)^{d-k}\left[\binom{n+2d-1}{d-k}-\binom{n+2d-1}{d-1-k}\right]\binom{n-1+k}{k}

which is the coefficient of tdt^{d} in the formal power series

[(1+t)n+2d1t(1+t)n+2d1]1(1+t)n=(1+t)2d1t(1+t)2d1\left[(1+t)^{n+2d-1}-t(1+t)^{n+2d-1}\right]\cdot\frac{1}{(1+t)^{n}}=(1+t)^{2d-1}-t(1+t)^{2d-1}

which equals

(2d1d)(2d1d1)={1,d=0,0,d>0.\binom{2d-1}{d}-\binom{2d-1}{d-1}=\begin{cases}1,&d=0,\\ 0,&d>0.\end{cases}

Now we proceed to the induction step. By Lemma 3.1 it is sufficient to show that the equalities in (1) hold after differentiation by =x1+x2++xn\ell=x_{1}+x_{2}+\dots+x_{n} and after restriction to Xn=0X_{n}=0.

We have that

An,d,m=(2m+1)k=0d(1)dk(n+2ddk)(n+2k)Sn,k,2m=(2m+1)(n+2d)Bn,d,m={22d(2d+1)!(n+2d)hn,2d,m=d0,m<d\begin{split}\ell\circ A_{n,d,m}&=(2m+1)\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}(n+2k)S_{n,k,2m}=(2m+1)(n+2d)B_{n,d,m}\\ &=\begin{cases}2^{2d}(2d+1)!(n+2d)h_{n,2d},&m=d\\ 0,&m<d\end{cases}\end{split}

where we have to assume by induction that the equality holds for Bn,d,mB_{n,d,m}. We have that

Bn,d,m=1(2m+1)(n+2d)2An,d,m=2mn+2dk=0d(1)dk(n+2ddk)(n+2k)2Sn,k,2m1.\ell\circ B_{n,d,m}=\frac{1}{(2m+1)(n+2d)}\ell^{2}\circ A_{n,d,m}=\frac{2m}{n+2d}\sum_{k=0}^{d}(-1)^{d-k}\binom{n+2d}{d-k}(n+2k)^{2}S_{n,k,2m-1}.

We subtract 2m(n+2d)An,d,m12m(n+2d)A_{n,d,m-1}, which by induction on mm is zero, from this and get

Bn,d,m=2mn+2dk=0d1(1)dk(n+2ddk)[(n+2k)2(n+2d)2]Sn,k,2m1=2mn+2dk=0d1(1)dk(n+2ddk)(2k2d)(2n+2k+2d)Sn,k,2m1=8m(n+2d1)k=0d1(1)d1k(n+2d2d1k)Sn,k,2m1=8m(n+2d1)An,d1,m1={8d(n+2d1)22d2(2d1)!hn,2d1m=d0m<d={(n+2d1)22d(2d)!hn,2d1m=d0m<d\begin{split}\ell\circ B_{n,d,m}&=\frac{2m}{n+2d}\sum_{k=0}^{d-1}(-1)^{d-k}\binom{n+2d}{d-k}\left[(n+2k)^{2}-(n+2d)^{2}\right]S_{n,k,2m-1}\\ &=\frac{2m}{n+2d}\sum_{k=0}^{d-1}(-1)^{d-k}\binom{n+2d}{d-k}(2k-2d)(2n+2k+2d)S_{n,k,2m-1}\\ &=8m(n+2d-1)\sum_{k=0}^{d-1}(-1)^{d-1-k}\binom{n+2d-2}{d-1-k}S_{n,k,2m-1}\\ &=8m(n+2d-1)A_{n,d-1,m-1}=\begin{cases}8d(n+2d-1)2^{2d-2}(2d-1)!h_{n,2d-1}&m=d\\ 0&m<d\end{cases}\\ &=\begin{cases}(n+2d-1)2^{2d}(2d)!h_{n,2d-1}&m=d\\ 0&m<d\end{cases}\end{split}

which agrees with the derivative of the right-hand side of (1).

Next, we restrict to Xn=0X_{n}=0 where we have Sn,k,d|Xn=0=i=0kSn1,i,dS_{n,k,d}|_{X_{n}=0}=\sum_{i=0}^{k}S_{n-1,i,d}. Hence

An,d,m|Xn=0=k=0d(1)dki=0k(n+2ddk)Sn1,i,2m+1==i=0dj=0di(1)j(n+2dj)Sn1,i,2m+1=i=0d(1)di(n1+2ddi)Sn1,i,2m+1=An1,d,m\begin{split}A_{n,d,m}|_{X_{n}=0}&=\sum_{k=0}^{d}(-1)^{d-k}\sum_{i=0}^{k}\binom{n+2d}{d-k}S_{n-1,i,2m+1}=\\ &=\sum_{i=0}^{d}\sum_{j=0}^{d-i}(-1)^{j}\binom{n+2d}{j}S_{n-1,i,2m+1}=\sum_{i=0}^{d}(-1)^{d-i}\binom{n-1+2d}{d-i}S_{n-1,i,2m+1}\\ &=A_{n-1,d,m}\end{split}

and

Bn,d,m|Xn=0=k=0d(1)dki=0k[(n+2d1dk)(n+2d1d1k)]Sn1,i,2m=i=0dj=0di(1)j[(n+2d1j)(n+2d1j1)]Sn1,i,2m=i=0dj=0di(1)di[(n+2d1di)(n+2d1di1)]Sn1,i,2m=Bn1,d,m.\begin{split}B_{n,d,m}|_{X_{n}=0}&=\sum_{k=0}^{d}(-1)^{d-k}\sum_{i=0}^{k}\left[\binom{n+2d-1}{d-k}-\binom{n+2d-1}{d-1-k}\right]S_{n-1,i,2m}\\ &=\sum_{i=0}^{d}\sum_{j=0}^{d-i}(-1)^{j}\left[\binom{n+2d-1}{j}-\binom{n+2d-1}{j-1}\right]S_{n-1,i,2m}\\ &=\sum_{i=0}^{d}\sum_{j=0}^{d-i}(-1)^{d-i}\left[\binom{n+2d-1}{d-i}-\binom{n+2d-1}{d-i-1}\right]S_{n-1,i,2m}=B_{n-1,d,m}.\end{split}

Remark 3.3.

For either even degree 2d2d or odd degree 2d+12d+1, our power sum decomposition in Theorem 3.2 has (n+dd)\binom{n+d}{d} terms. This gives an upper bound for the Waring rank of the complete symmetric forms. According to the Alexander–Hirschowitz Theorem, except in a few well-understood cases [AH, I], general forms of degree 2d+12d+1 have Waring rank 1n(n+2d2d+1)\lceil\frac{1}{n}\binom{n+2d}{2d+1}\rceil and general forms of degree 2d2d have Waring rank 1n(n+2d12d)\lceil\frac{1}{n}\binom{n+2d-1}{2d}\rceil. Observe that our bound has degree dd as a polynomial in nn while the Waring rank for general forms grows like a polynomial of degree 2d2d or 2d12d-1 in nn. We believe that our bound is sharp for odd degree (i.e. that we are actually giving a Waring decomposition), and will show it for d=1d=1 (i.e. degree 3) in the next section. For even degree, experimentally we have seen that it is not necessarily sharp.

In fact, for all n14n\neq 14, we can find coefficients α0,α1,α2,α3,α4\alpha_{0},\alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4} so that

hn,4=α0h14+i=1n(α1h1+α2Xi)4+1i<jn(α3h1+α4Xi+α4Xj)4.h_{n,4}=\alpha_{0}h_{1}^{4}+\sum_{i=1}^{n}(\alpha_{1}h_{1}+\alpha_{2}X_{i})^{4}+\sum_{1\leq i<j\leq n}(\alpha_{3}h_{1}+\alpha_{4}X_{i}+\alpha_{4}X_{j})^{4}.

This expression has (n+22)n\binom{n+2}{2}-n terms which is nn fewer terms than the expression from Theorem. 3.2. In particular, for n=13n=13 we can get explicit values for the coefficients and we can write

h13,4=23h14+124i=113(h1+Xi)4+1241i<j13(Xi+Xj)4.h_{13,4}=-\frac{2}{3}h_{1}^{4}+\frac{1}{24}\sum_{i=1}^{13}(h_{1}+X_{i})^{4}+\frac{1}{24}\sum_{1\leq i<j\leq 13}(X_{i}+X_{j})^{4}.

We will look more into this at the end of Section 5.

4. Symmetric cubic polynomials

In this section we deeply analyze the case of symmetric cubic polynomials. We begin by considering the Waring rank. In the case of degree three, Theorem 3.2 applied with d=1d=1 states that

h3=124(i=1n(h1+2Xi)3(n+2)h13)\displaystyle h_{3}=\frac{1}{24}\left(\sum_{i=1}^{n}(h_{1}+2X_{i})^{3}-(n+2)h_{1}^{3}\right)

and we will show that this is indeed a Waring decomposition, not just for this particular symmetric cubic, but that a general symmetric cubic has a similar Waring decomposition.

We will study the artinian Gorenstein algebras defined by the annihilators of symmetric cubic forms in n3n\geq 3 variables. We will determine the possible Hilbert function of the annihilator, the Waring rank, the cactus rank, and the resolution of all such Gorenstein algebras. We will also show that they all satisfy the SLP and that there are three linear forms that are sufficient to provide Lefschetz elements for them all.

We will parameterize the symmetric forms of degree three by the projective plane using the basis {p13,np1p2,n2p3}\{p_{1}^{3},np_{1}p_{2},n^{2}p_{3}\}, where pi=j=1nXjip_{i}=\sum_{j=1}^{n}X_{j}^{i} are the power sum symmetric forms in E=k[X1,X2,,Xn]E=k[X_{1},X_{2},\dots,X_{n}]. Let k[a0,a1,a2]k[a_{0},a_{1},a_{2}] be the coordinate ring of this plane corresponding to the given basis.

We start by defining a rational cubic curve as a set of symmetric cubics whose Waring rank is at most nn. For (α:β)(\alpha\colon\beta) in 1\mathbb{P}^{1}, define (up to a scalar multiple) a symmetric cubic

(2) F(α,β)\displaystyle F(\alpha,\beta) =i=1n(αnXi+βp1)3\displaystyle=\sum_{i=1}^{n}(\alpha nX_{i}+\beta p_{1})^{3}
=α3n3p3+3α2βn2p1p2+(3αβ2+β3)np13.\displaystyle=\alpha^{3}n^{3}p_{3}+3\alpha^{2}\beta n^{2}p_{1}p_{2}+(3\alpha\beta^{2}+\beta^{3})np_{1}^{3}.

Furthermore, let 𝒞\mathcal{C} be the rational cubic curve defined as the image of the map

γ:12=(p13,np1p2,n2p3),(α:β)(3αβ2+β3:3α2β:α3).\gamma\colon\mathbb{P}^{1}\to\mathbb{P}^{2}=\mathbb{P}(\langle p_{1}^{3},np_{1}p_{2},n^{2}p_{3}\rangle),\ (\alpha\colon\beta)\mapsto(3\alpha\beta^{2}+\beta^{3}:3\alpha^{2}\beta:\alpha^{3}).

(Note the coefficients nn and n2n^{2} on the basis.) That means, the image of (α:β)(\alpha\colon\beta) corresponds to the cubic F(α,β)F(\alpha,\beta).

We also will use the above coordinates to parameterize any symmetric cubic in nn variables. In fact, since the power sums whose degrees are at most nn generate the ring of symmetric polynomials, every symmetric cubic can be written (over a field of characteristic zero) as F=a0p13+a1np1p2+a2n2p3F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3}. We refer to FF as the cubic at the point (a0:a1:a2)(a_{0}\colon a_{1}\colon a_{2}).

Notation 4.1.

We fix three lines in 2\mathbb{P}^{2} which will play an important role in this section:

1\ell_{1} is the line defined by a1+3a2=0a_{1}+3a_{2}=0,

2\ell_{2} is the line defined by a2=0a_{2}=0, and

3\ell_{3} is the line defined by a0+a1+a2=0a_{0}+a_{1}+a_{2}=0.

Lemma 4.2.

The cubic curve 𝒞\mathcal{C} is a cuspidal curve given by the equation

27a0a229a12a2a13=0.27a_{0}a_{2}^{2}-9a_{1}^{2}a_{2}-a_{1}^{3}=0.

Its cusp is 𝒫=(1:0:0)\mathcal{P}=(1:0:0) and its unique flex point is 𝒬=(2:3:1)\mathcal{Q}=(2:-3:1). The line through 𝒫\mathcal{P} and 𝒬\mathcal{Q} is the line 1\ell_{1}. Moreover, 2\ell_{2} is the unique line meeting 𝒞\mathcal{C} at 𝒫\mathcal{P} with multiplicity three.

Proof.

It is elementary to check that 𝒞\mathcal{C} has a singularity at (1:0:0)(1:0:0). This is a cusp rather than a node because there is only one point (α:β)(\alpha:\beta) in 1\mathbb{P}^{1} such that γ((α:β))=(1:0:0)\gamma((\alpha:\beta))=(1:0:0), namely (0:1)(0:1). The fact that a cuspidal cubic has only one flex point can be found in [BCGM, page 146]. One can check that the tangent line at 𝒬=(2:3:1)\mathcal{Q}=(2:-3:1) is the line 3\ell_{3} and that the intersection of this line with 𝒞\mathcal{C} at 𝒬\mathcal{Q} has multiplicity 3. Hence 𝒬\mathcal{Q} is indeed a flex point. The rest is immediate. ∎

Below we slightly abuse notation and refer to the line 2\ell_{2} as the tangent line to 𝒞\mathcal{C} at 𝒫\mathcal{P}.

Proposition 4.3.

The Hilbert function of S/ann(a0p13+a1np1p1+a2n2p3)S/\operatorname{ann}(a_{0}p_{1}^{3}+a_{1}np_{1}p_{1}+a_{2}n^{2}p_{3}) is

{(1,1,1,1)at 𝒫=(1:0:0)(1,n1,n1,1)at 𝒬=(2,3,1)(1,n,n,1)otherwise.\begin{cases}(1,1,1,1)&\text{at $\mathcal{P}=(1:0:0)$}\\ (1,n-1,n-1,1)&\text{at $\mathcal{Q}=(2,-3,1)$}\\ (1,n,n,1)&\text{otherwise.}\end{cases}
Proof.

Since ann(F)\operatorname{ann}(F) is 𝔖n\mathfrak{S}_{n}-invariant there are three possibilities for [ann(F)]1[S]1=V0V1[\operatorname{ann}(F)]_{1}\subseteq[S]_{1}=V_{0}\oplus V_{1}, where V0V_{0} is the trivial representation and V1V_{1} is the standard representation. If [ann(F)]10[\operatorname{ann}(F)]_{1}\neq 0, we have [ann(F)]1=V0[\operatorname{ann}(F)]_{1}=V_{0} or [ann(F)]1=V1[\operatorname{ann}(F)]_{1}=V_{1}. In the second case, [ann(F)]1=p1[\operatorname{ann}(F)]_{1}=\langle p_{1}\rangle^{\perp} and F=λp13F=\lambda p_{1}^{3} and in the first case [ann(F)]1=[p1]1[\operatorname{ann}(F)]_{1}=[\langle p_{1}\rangle]_{1}. There is a unique solution to p1F=0p_{1}\circ F=0 since

p1(a0p13+a1np1p2+a2n2p3)=3na0p12+a1n2p2+2na1p12+3a2n2p2=(3a0+2a1)np12+(a1+3a2)n2p2.\begin{split}p_{1}\circ(a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3})=3na_{0}p_{1}^{2}+a_{1}n^{2}p_{2}+2na_{1}p_{1}^{2}+3a_{2}n^{2}p_{2}\\ =(3a_{0}+2a_{1})np_{1}^{2}+(a_{1}+3a_{2})n^{2}p_{2}.\end{split}

In all other points [ann(F)]1=0[\operatorname{ann}(F)]_{1}=0. ∎

In order to find the Waring rank of all symmetric cubic forms we need the following lemma about small orbits in n1\mathbb{P}^{n-1} under the action of 𝔖n\mathfrak{S}_{n}. We will assume that n3n\geq 3.

Lemma 4.4.

If XX is an orbit of points in n1\mathbb{P}^{n-1} under the action of 𝔖n\mathfrak{S}_{n} with |X|n|X|\leq n, we have that one of the following holds

  1. (1)

    |X|=1|X|=1 and X={(1:1::1)}X=\{(1:1:\cdots:1)\}.

  2. (2)

    |X|=n|X|=n and X={(α:β::β)}X=\{(\alpha:\beta:\cdots:\beta)\}, αβ\alpha\neq\beta.

  3. (3)

    n=3n=3 and X={(1:ξ:ξ2),(1:ξ2:ξ)}X=\{(1:\xi:\xi^{2}),(1:\xi^{2}:\xi)\}, where ξ2+ξ+1=0\xi^{2}+\xi+1=0.

  4. (4)

    n=4n=4 and X={(1:1:1:1),(1:1:1:1),(1:1:1:1)}X=\{(1:1:-1:-1),(1:-1:1:-1),(1:-1:-1:1)\}.

Proof.

If |X|>1|X|>1, we have at least one point of the form (1:t1::tn1)(1:t_{1}:\dots:t_{n-1}). If not all tit_{i} are equal, we have at least n1n-1 points of this form. In the case n=3n=3, this gives a possible orbit {(1:t1:t2),(1:t2:t1)}\{(1:t_{1}:t_{2}),(1:t_{2}:t_{1})\} if (t2:t1:1)=(1:t1:t2)(t_{2}:t_{1}:1)=(1:t_{1}:t_{2}) or (t2:t1:1)=(1:t2:t1)(t_{2}:t_{1}:1)=(1:t_{2}:t_{1}). In the first case, t2=1t_{2}=1 and we get |X|=3|X|=3 or |X|=1|X|=1. In the second case, we get t22=t1t_{2}^{2}=t_{1}, t12=t2t_{1}^{2}=t_{2}, so t13=t23=1t_{1}^{3}=t_{2}^{3}=1. Thus we are we are either in case (1) or case (3). When n=4n=4, we get a possible orbit if X={(1:t1:t1:t2),(1:t2:t1:t2),(1:t2:t2:t1)}X=\{(1:t_{1}:t_{1}:t_{2}),(1:t_{2}:t_{1}:t_{2}),(1:t_{2}:t_{2}:t_{1})\}. If t1t2t_{1}\neq t_{2} we must have (t1:1:t1:t2)=(1:t2:t1:t2)(t_{1}:1:t_{1}:t_{2})=(1:t_{2}:t_{1}:t_{2}) or (t1:1:t1:t2)=(1:t2:t2:t1)(t_{1}:1:t_{1}:t_{2})=(1:t_{2}:t_{2}:t_{1}). In the first case we get t1=1t_{1}=1 and we are in case (2). In the second case we get t12=t22t_{1}^{2}=t_{2}^{2} and t1t2=1t_{1}t_{2}=1 which gives case (4). If n5n\geq 5, we have to have (1:t1:t2::t2)(1:t_{1}:t_{2}:\cdots:t_{2}) and t2=1t_{2}=1 which leads to the case (2). ∎

Proposition 4.5.

The Waring rank of the symmetric cubic form F=a0p13+a1np1p2+a2n2p3F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3} at (a0:aa:a2)(a_{0}:a_{a}:a_{2}) is

wr(F)={1at 𝒫=(1:0:0);n1at 𝒬=(2:3:1) if n=3nat 𝒞{𝒫} and, if n=3, also at 1{𝒬};2(n1)at 2{𝒫};n+1otherwise.wr(F)=\begin{cases}1&\text{at $\mathcal{P}=(1:0:0)$;}\\ n-1&\text{at $\mathcal{Q}=(2:-3:1)$ if $n=3$; }\\ n&\text{at $\mathcal{C}\setminus\{\mathcal{P}\}$ and, if $n=3$, also at $\ell_{1}\setminus\{\mathcal{Q}\}$};\\ 2(n-1)&\text{at $\ell_{2}\setminus\{\mathcal{P}\}$};\\ n+1&otherwise.\end{cases}
Proof.

The Waring rank is 11 if and only if the Hilbert function is (1,1,1,1)(1,1,1,1) and we are at (1:0:0)(1:0:0) which is the cusp of 𝒞\mathcal{C} according to Lemma 4.2.

We have that the Waring rank has to be at least n1n-1 at the flex point (2:3:1)(2:-3:1) and at least nn at all other points. If the rank is n1n-1 at the flex point we would have [ann(F)]2=[IX]2[\operatorname{ann}(F)]_{2}=[I_{X}]_{2} where XX is the support of the Waring decomposition. Since FF is symmetric, XX has to be a union of orbits and Lemma 4.4 says that we can only have n=3n=3 or n=4n=4. In the case n=3n=3, we do get a Waring decomposition of FF as

F=(X1+ξX2+ξ2X3)3+(X1+ξ2X2+ξX3)3=2p139p1p2+9p3.F=(X_{1}+\xi X_{2}+\xi^{2}X_{3})^{3}+(X_{1}+\xi^{2}X_{2}+\xi X_{3})^{3}=2p_{1}^{3}-9p_{1}p_{2}+9p_{3}.

In the case n=4n=4, there is no non-zero symmetric cubic that can be expressed in terms of the three linear forms X1+X2X3X4,X1X2+X3X4,X1X2X3+X4X_{1}+X_{2}-X_{3}-X_{4},X_{1}-X_{2}+X_{3}-X_{4},X_{1}-X_{2}-X_{3}+X_{4}. Thus, for n4n\geq 4, the Waring rank has to be nn at the flex point.

For points on the curve 𝒞\mathcal{C} we have a power sum decomposition with nn terms (see Equation (2)). Moreover, for all these points apart from the flex point and the cusp, the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,n,n,1)(1,n,n,1) by Proposition 4.3. Hence the Waring rank of FF has to be equal to nn.

For points on the line a2=0a_{2}=0, which is the tangent line to 𝒞\mathcal{C} at the cusp, we have that F=a0p13+a1np1p2F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}. Let Y={(λixi)2:i=1nλi=i=1nλi2}Y=\{(\sum\lambda_{i}x_{i})^{2}\colon\sum_{i=1}^{n}\lambda_{i}=\sum_{i=1}^{n}\lambda_{i}^{2}\}. We have that Yann(F)Y\subseteq\operatorname{ann}(F) and dimY=n2\dim Y=n-2 as an affine variety. Let V=[ann(F)]2V=[\operatorname{ann}(F)]_{2}. If FF has a Waring decomposition corresponding to the reduced set of points XX, we have that [IX]2Y={0}[I_{X}]_{2}\cap Y=\{0\}. We hence have that

dim(Y)+dim([IX]2)<dim([ann(F)]2)\dim\mathbb{P}(Y)+\dim\mathbb{P}([I_{X}]_{2})<\dim\mathbb{P}([\operatorname{ann}(F)]_{2})

which gives

HX(2)>n+n3=2n3.H_{X}(2)>n+n-3=2n-3.

This shows that wr(F)2(n1)wr(F)\geq 2(n-1). We will now show that FF has a Waring decomposition of length 2(n1)2(n-1). Let R=k[y1,y2,,yn1]=k[x1xn,x2xn,,xn1xn]SR=k[y_{1},y_{2},\dots,y_{n-1}]=k[x_{1}-x_{n},x_{2}-x_{n},\dots,x_{n-1}-x_{n}]\subseteq S and let IZI_{Z} denote the ideal in SS generated by (xixj)(xi+xj2xk)(x_{i}-x_{j})(x_{i}+x_{j}-2x_{k}), for ik,jki\neq k,j\neq k. The ideal JJ in RR generated by the same forms is an artinian Gorenstein ideal given by the annihilator of h2(Y1,Y2,,Yn)h_{2}(Y_{1},Y_{2},\dots,Y_{n}). Thus ZZ is a zero-dimensional Gorenstein scheme of length n+1n+1 concentrated at the point (1:1::1)(1:1:\cdots:1). Moreover, it is contained in ann(F)\operatorname{ann}(F). We now take a Waring decomposition of h2(Y1,Y2,,Yn)h_{2}(Y_{1},Y_{2},\dots,Y_{n}) which has length n1n-1 since h2h_{2} is a non-degenerate quadric in n1n-1 variables. This gives a reduced subscheme of length n1n-1 and the cone over this is a reduced curve 𝒟\mathcal{D} consisting of n1n-1 lines meeting at the point (1:1::1)(1:1:\cdots:1). The ideal ann(F)\operatorname{ann}(F) is artinian and contains I𝒟I_{\mathcal{D}}. Hence Bertini’s theorem says that we can find a quadric in ann(F)\operatorname{ann}(F) intersecting 𝒟\mathcal{D} in 2(n1)2(n-1) distinct points. The ideal of these points is contained in ann(F)\operatorname{ann}(F) which shows that wr(F)2(n1)wr(F)\leq 2(n-1).

It remains to show that wr(F)=n+1wr(F)=n+1 away from 𝒞\mathcal{C} and its tangent line at the cusp. Through any such point, we have a line through the cusp, meeting the curve 𝒞\mathcal{C} in exactly one more point. This gives a decomposition into n+1n+1 cubes of linear forms. Since the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,n,n,1)(1,n,n,1), wr(F)nwr(F)\geq n and if it is equal to nn, the ideal of the corresponding points equals the ideal ann(F)\operatorname{ann}(F) in degree 22. Hence the points are invariant under the action of 𝔖n\mathfrak{S}_{n} and by Lemma 4.4 the only possible orbits give points on the curve 𝒞\mathcal{C}. ∎

Refer to caption
Figure 1. The cubic curve and the lines

Before looking at the cactus rank, we will study the structure of the resolutions of S/ann(F)S/\operatorname{ann}(F) for symmetric cubics.

Proposition 4.6.

For any symmetric cubic F=a0p13+a1np1p2+a2n2p3F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3} the Betti numbers of S/ann(F)S/\operatorname{ann}(F) can be obtained from the Betti numbers of the coordinate ring of a zero-dimensional locally Gorenstein scheme XX. We can write ann(F)=IX+JX\operatorname{ann}(F)=I_{X}+J_{X}, where JXJ_{X} is a canonical ideal and for i=0,1,,ni=0,1,\dots,n and all jj we have that

βi,j(S/ann(F))=βi,j(S/IX)+βni,3+nj(S/IX).\beta_{i,j}(S/\operatorname{ann}(F))=\beta_{i,j}(S/I_{X})+\beta_{n-i,3+n-j}(S/I_{X}).

There are the following cases for the resolution of IXI_{X}.

  • (i)(i)

    For FF at 𝒫\mathcal{P}, XX is a single point and the resolution of S/IXS/I_{X} is linear with βi,i=(n1i)\beta_{i,i}=\binom{n-1}{i}, for i=0,1,,n1i=0,1,\dots,n-1.

  • (ii)(ii)

    For FF at 𝒬\mathcal{Q}, and n=3n=3, XX is two points and the resolution of S/IXS/I_{X} is a Koszul complex with β0,0=β1,1=β1,2=β2,4=1\beta_{0,0}=\beta_{1,1}=\beta_{1,2}=\beta_{2,4}=1.

  • (iii)(iii)

    For FF at 𝒬\mathcal{Q}, and n>3n>3, XX is an arithmetically Gorenstein set of nn points in a hyperplane and the resolution is forced by the hh-vector (1,n2,1)(1,n-2,1), i.e., β0,0=β1,1=βn2,n=βn1,n+1=1\beta_{0,0}=\beta_{1,1}=\beta_{n-2,n}=\beta_{n-1,n+1}=1 and

    βi,i+1=i(n2i)n1(ni+1)+(i1)(n1i)n1(ni),i=1,2,,n2.\beta_{i,i+1}=\frac{i(n-2-i)}{n-1}\binom{n}{i+1}+\frac{(i-1)(n-1-i)}{n-1}\binom{n}{i},\quad i=1,2,\dots,n-2.
  • (iv)(iv)

    For FF on 1{𝒫,𝒬\ell_{1}\setminus\{\mathcal{P},\mathcal{Q}}, and n=3n=3, XX is a set of three points not on a line, with a linear resolution, i.e., β0,0=1\beta_{0,0}=1, β1,2=3\beta_{1,2}=3 and β2,3=2\beta_{2,3}=2.

  • (v)(v)

    For FF on 1{𝒫,𝒬}\ell_{1}\setminus\{\mathcal{P},\mathcal{Q}\}, and n>3n>3, XX is a set of n+1n+1 points where nn of them are linearly general in a hyperplane. The resolution of S/IXS/I_{X} is linear except for the last two homological degrees where βn2,n=βn1,n+1=1\beta_{n-2,n}=\beta_{n-1,n+1}=1. The Betti numbers are

    βi,i+1=i(n1i)n(n+1i+1),i=1,2,,n2.\beta_{i,i+1}=\frac{i(n-1-i)}{n}\binom{n+1}{i+1},\qquad i=1,2,\dots,n-2.
  • (vi)(vi)

    For FF at 𝒞\mathcal{C} away from 𝒫\mathcal{P} and 𝒬\mathcal{Q}, XX is a set of nn points not in a hyperplane. The resolution of S/IXS/I_{X} is linear with β0,0=1\beta_{0,0}=1 and

    βi,i+1=i(ni+1),i=1,2,,n1.\beta_{i,i+1}=i\binom{n}{i+1},\qquad i=1,2,\dots,n-1.
  • (vii)(vii)

    For FF on 2{𝒫}\ell_{2}\setminus\{\mathcal{P}\}, XX is a non-degenerate arithmetically Gorenstein scheme of length n+1n+1 concentrated at a single point and S/IXS/I_{X} has an almost linear resolution, i.e., β0,0=βn1,n+1=1\beta_{0,0}=\beta_{n-1,n+1}=1 and

    βi,i+1=i(n1i)n(n+1i+1),i=1,2,,n2.\beta_{i,i+1}=\frac{i(n-1-i)}{n}\binom{n+1}{i+1},\qquad i=1,2,\dots,n-2.
  • (viii)(viii)

    For all other FF, XX is an arithmetically Gorenstein set of n+1n+1 points and S/IXS/I_{X} has an almost linear resolution, i.e., β0,0=βn1,n+1=1\beta_{0,0}=\beta_{n-1,n+1}=1 and

    βi,i+1=i(n1i)n(n+1i+1),i=1,2,,n2.\beta_{i,i+1}=\frac{i(n-1-i)}{n}\binom{n+1}{i+1},\qquad i=1,2,\dots,n-2.
Proof.

The existence of the set XX comes from the Waring decompositions of Proposition 4.5 except for in case (vii)(vii) and we will start by looking at what happens in that case. Here we have F=a0p13+a1np1p2F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2} for some coefficients a0,a1a_{0},a_{1}, where a10a_{1}\neq 0. Let JJ be the ideal generated by the quadratic polynomials in the subring R=k[x1x2,x1x3,,x1xn]R=k[x_{1}-x_{2},x_{1}-x_{3},\dots,x_{1}-x_{n}] that annihilate p2p_{2}. All such generators annihilate p13p_{1}^{3} and p1p2p_{1}p_{2}, so Jann(F)J\subseteq\operatorname{ann}(F) for any F=a0p13+a1np1p2F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}. Furthermore, A=R/(RJ)A=R/(R\cap J) is artinian with Hilbert function (1,n1,1)(1,n-1,1). It is Gorenstein, since there is no socle in degree one. Indeed, (αixi)(βixi)p2=2αiβi(\sum\alpha_{i}x_{i})(\sum\beta_{i}x_{i})\circ p_{2}=2\sum\alpha_{i}\beta_{i} which gives a non-degenerate pairing on R1R_{1}. Thus we have that J=IXJ=I_{X} where XX is a zero-dimensional arithmetically Gorenstein subscheme concentrated at the point (1:1::1)(1:1:\cdots:1). In fact, AA is the subring of S/ann(p2)S/\operatorname{ann}(p_{2}) generated by x1x2,x1x3,,x1xnx_{1}-x_{2},x_{1}-x_{3},\dots,x_{1}-x_{n}. The symmetric quadric qq in ann(F)\operatorname{ann}(F) is given by

q=(3a0+(n+2)a1)p123n(a0+a1)p2q=(3a_{0}+(n+2)a_{1})p_{1}^{2}-3n(a_{0}+a_{1})p_{2}

and evaluation at (1,1,,1)(1,1,\dots,1) gives

q(1,1,,1)=(3a0+(n+2)a1)n23n(a0+a1)n=a1n2(n1)0q(1,1,\dots,1)=(3a_{0}+(n+2)a_{1})n^{2}-3n(a_{0}+a_{1})n=a_{1}n^{2}(n-1)\neq 0

showing that qq is a non-zero-divisor on XX. Since XX is arithmetically Gorenstein, the ideal (q)(q) is a canonical ideal and we have ann(F)=IX+(q)\operatorname{ann}(F)=I_{X}+(q).

In all the other cases, the Castelnuovo–Mumford regularity of the set of points XX is at most 22 and we can apply [B, Theorem 3.4] to conclude that the image of the ideal ann(F)\operatorname{ann}(F) in the coordinate ring of XX is a canonical ideal. When the Castelnuovo–Mumford regularity of XX is at most one, the statement about the Betti numbers follows from [B, Proposition 3.5]. In the cases where the Castelnuovo–Mumford regularity is two and the scheme XX is arithmetically Gorenstein, the canonical ideal is principal, which means that it is generated by a non-zero-divisor. Hence the statement about the Betti numbers holds also in this case.

In case (vi)(vi), S/ann(F)S/\operatorname{ann}(F) is the quotient of the coordinate ring of XX by a non-zero-divisor and the ideal is still a canonical ideal since XX is arithmetically Gorenstein.

It remains to treat the case (v)(v) where the Castelnuovo–Mumford regularity of XX is two but the ideal is not given by a non-zero-divisor. We look at the short exact sequence

0JXS/IXS/(IX+JX)00\longrightarrow J_{X}\longrightarrow S/I_{X}\longrightarrow S/(I_{X}+J_{X})\longrightarrow 0

where JXJ_{X} is the canonical ideal that is the image of ann(F)\operatorname{ann}(F) in S/IXS/I_{X}. Since there may be cancellations, we only get inequalities

βi,j(S/ann(F))βi,j(S/IX)+βni,3+nj(S/IX)\beta_{i,j}(S/\operatorname{ann}(F))\leq\beta_{i,j}(S/I_{X})+\beta_{n-i,3+n-j}(S/I_{X})

but in our case, the only possible cancellation would be in degree 33 and dually in degree nn. If this cancellation occurred, it would mean that the ideal ann(F)\operatorname{ann}(F) was generated by quadrics, but this is not possible, since the symmetric quadric qq is not a non-zero-divisor on XX since it vanishes at the point (1:1::1)(1:1:\cdots:1).

For each of the cases, we will now prove the statements about the resolutions of S/IXS/I_{X}.

  • (i)(i)

    For F=p13F=p_{1}^{3}, XX is the single point (1:1::1)(1:1:\cdots:1) and the ideal is generated by the linear forms x1x2,x1x3,,x1xnx_{1}-x_{2},x_{1}-x_{3},\dots,x_{1}-x_{n}. Hence the resolution is linear and given by a Koszul complex of length n1n-1, which gives the stated Betti numbers.

  • (ii)(ii)

    Here the Waring rank is two and we have that the ideal of two points in 2\mathbb{P}^{2} is a complete intersection of type (1,2)(1,2). The resolution is a Koszul complex which gives the stated Betti numbers.

  • (iii)(iii)

    The Waring rank is nn and the set of points giving the Waring decomposition is the 𝔖n\mathfrak{S}_{n}-orbit of (1n:1:1::1)(1-n:1:1:\cdots:1). They lie in a hyperplane and are in linearly general position in that hyperplane. Hence they form an arithmetically Gorenstein set of points with hh-vector (1,n2,1)(1,n-2,1). The Betti numbers of the coordinate ring S/IXS/I_{X} are given as sums of the consecutive Betti numbers of an artinian Gorenstein algebra with the same hh-vector. This means that β0,0=β1,1=βn2,n=βn1,n+1=1\beta_{0,0}=\beta_{1,1}=\beta_{n-2,n}=\beta_{n-1,n+1}=1 and

    βi,i+1=i(n2i)n1(ni+1)+(i1)(n1i)n1(ni),i=1,2,,n2,\beta_{i,i+1}=\frac{i(n-2-i)}{n-1}\binom{n}{i+1}+\frac{(i-1)(n-1-i)}{n-1}\binom{n}{i},\quad i=1,2,\dots,n-2,

    where we get the Betti numbers for the artinian Gorenstein algebra with hh-vector (1,n2,1)(1,n-2,1) by the formula for the Betti numbers of a pure resolution ([HK]).

  • (iv)(iv)

    We have that X={(1:1:1),(1:ξ:ξ2),(1:ξ2,ξ)}X=\{(1:1:1),(1:\xi:\xi^{2}),(1:\xi^{2},\xi)\}, where ξ2+ξ+1=0\xi^{2}+\xi+1=0 and the three points do not lie on a line. For three points not on a line in 2\mathbb{P}^{2}, the resolution is linear and the Betti numbers are β0,0=1\beta_{0,0}=1, β1,2=3\beta_{1,2}=3 and β2,3=2\beta_{2,3}=2.

  • (v)(v)

    XX is the union of the single point orbit X1={(1:1::1)}X_{1}=\{(1:1:\cdots:1)\} and the nn-point orbit X2X_{2} of (1n:1:1::1)(1-n:1:1:\cdots:1). Since X2X_{2} lies in a hyperplane, but is in linearly general position within that hyperplane, X2X_{2} is arithmetically Gorenstein with hh-vector (1,n2,1)(1,n-2,1). Thus the hh-vector of XX has to be (1,n1,1)(1,n-1,1). For an artinian reduction B=S/JB=S/J of S/IXS/I_{X}, the ideal in degree 22 equals the ideal of an artinian reduction of S/IX2S/I_{X_{2}}. Thus we see that BB has a one-dimensional socle in degree 22 and a one-dimensional socle in degree 11. This gives us the last column in the Betti table of BB and forces that βn2,2=1\beta_{n-2,2}=1. By duality, we get that βi,i+2=0\beta_{i,i+2}=0 for i=0,1,,n3i=0,1,\dots,n-3. Hence all the remaining Betti numbers are equal to the Betti numbers of an arithmetically Gorenstein scheme with the same hh-vector as XX, i.e., the same as in case (vii)(vii) below.

  • (vi)(vi)

    Here XX is the 𝔖n\mathfrak{S}_{n}-orbit of (nα+β:β::β)(n\alpha+\beta:\beta:\cdots:\beta) where α0\alpha\neq 0 and α+β0\alpha+\beta\neq 0. Thus XX is a set of nn points not in a hyperplane with hh-vector (1,n1)(1,n-1) showing that the resolution is linear. Again, we can use the formula for the Betti numbers of a pure resolution to write

    βi,i+1=i(ni+1),i=1,2,,n1.\beta_{i,i+1}=i\binom{n}{i+1},\qquad i=1,2,\dots,n-1.
  • (vii)(vii)

    As we have seen, XX is an arithmetically Gorenstein subscheme of length n+1n+1 concentrated at the point (1:1::1)(1:1:\cdots:1). The hh-vector is (1,n1,1)(1,n-1,1) and the resolution of S/IXS/I_{X} is almost linear since the artinian reduction is extremely compressed ([FL]). From the formula for the Betti numbers of a pure resolution, we get

    βi,i+1=i(n1i)n(n+1i+1),i=1,2,,n2.\beta_{i,i+1}=\frac{i(n-1-i)}{n}\binom{n+1}{i+1},\qquad i=1,2,\dots,n-2.
  • (viii)(viii)

    Here XX is the union of the single point orbit X1={(1:1::1)}X_{1}=\{(1:1:\cdots:1)\} and the nn-point orbit of {(nα+β:β::β)\{(n\alpha+\beta:\beta:\cdots:\beta) where α0\alpha\neq 0 and α+β0\alpha+\beta\neq 0. This set of points is in linearly general position with hh-vector (1,n1,1)(1,n-1,1). Thus XX is arithmetically Gorenstein and S/IXS/I_{X} has an almost linear resolution as in the case (vii)(vii) above.

We now turn to the cactus rank and we will also use the following result by K. Ranestad and F. -O. Schreyer.

Proposition 4.7 ([RS, Corollary 1]).

Let FF be a homogeneous polynomial. If the degree of any minimal generator of ann(F)\operatorname{ann}(F) is at most dd, then one has

cr(F)1ddeg(ann(F)).cr(F)\geq\frac{1}{d}\cdot\deg(\operatorname{ann}(F)).
Proposition 4.8.

The cactus rank of the symmetric cubic form F=a0p13+a1np1p2+a2n2p3F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3} at (a0:aa:a2)(a_{0}:a_{a}:a_{2}) is

cr(F)={1at 𝒫=(1:0:0);n1at 𝒬=(2:3:1), provided n=3;nat 𝒞{𝒫,𝒬} and, if n=3, also along 1{𝒬};n+1otherwise.cr(F)=\begin{cases}1&\text{at $\mathcal{P}=(1:0:0)$};\\ n-1&\text{at $\mathcal{Q}=(2:-3:1)$, provided $n=3$};\\ n&\text{at $\mathcal{C}\setminus\{\mathcal{P},\mathcal{Q}\}$ and, if $n=3$, also along $\ell_{1}\setminus\{\mathcal{Q}\}$};\\ n+1&otherwise.\end{cases}
Proof.

We will combine the lower bound provided by Proposition 4.7 and the upper bound cr(F)wr(F)cr(F)\leq wr(F), using that we know the Waring rank by Proposition 4.5. We consider cases depending on the Hilbert function of S/ann(F)S/\operatorname{ann}(F).

By Proposition 4.5, we know wr(F)=1wr(F)=1 if and only if FF is the form at 𝒫\mathcal{P}, i.e., F=p13F=p_{1}^{3}, which has cactus rank one.

Now consider FF at the flex point 𝒬\mathcal{Q} of 𝒞\mathcal{C}. If n=3n=3 then ann(F)\operatorname{ann}(F) is a complete intersection of type (1,2,3)(1,2,3) by Proposition 4.6(ii). Hence, this ideal does not contain the ideal of a point, which implies cr(F)2cr(F)\geq 2. Since wr(F)=2wr(F)=2 by Proposition 4.5, we get cr(F)=2cr(F)=2. If n4n\geq 4, then ann(F)\operatorname{ann}(F) is generated by quadrics (see Proposition 4.6(iii)) and has degree 2n+22n+2 by Proposition 4.3. Hence Proposition 4.7 gives cr(F)2n+22=n+1cr(F)\geq\frac{2n+2}{2}=n+1. We get equality because Proposition 4.5 yields wr(F)=n+1wr(F)=n+1.

It remains to consider cubics at points other than the cusp 𝒫\mathcal{P} and the flex point 𝒬\mathcal{Q}. Hence, the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,n,n,1)(1,n,n,1) by Proposition 4.3. In particular, ann(F)\operatorname{ann}(F) has degree 2n+22n+2. Observe that ann(F)\operatorname{ann}(F) is generated by quadrics, unless FF corresponds to a point on the curve 𝒞\mathcal{C} or a point on the line 1\ell_{1}. Excluding the latter two cases, Proposition 4.7 gives cr(F)2n+22=n+1cr(F)\geq\frac{2n+2}{2}=n+1. We get equality if FF is also not on the line 2\ell_{2} because then wr(F)=n+1wr(F)=n+1 by Proposition 4.5. Thus, we are left with considering three situations.

First, assume that FF is a point on (1𝒞){𝒫,𝒬}(\ell_{1}\cup\mathcal{C})\setminus\{\mathcal{P},\mathcal{Q}\}. If Zn1Z\subset\mathbb{P}^{n-1} is a zero-dimensional subscheme such that its ideal IZI_{Z} is contained in ann(F)\operatorname{ann}(F), then we get dimK[S/IZ]2dimK[S/ann(F)]2=n\dim_{K}[S/I_{Z}]_{2}\geq\dim_{K}[S/\operatorname{ann}(F)]_{2}=n, where the equality is due to Proposition 4.3. We conclude that degZdimK[S/IZ]2n\deg Z\geq\dim_{K}[S/I_{Z}]_{2}\geq n, which implies cr(F)ncr(F)\geq n. Unless FF is a point on the line 1\ell_{1} and n4n\geq 4, we have wr(F)=nwr(F)=n by Proposition 4.5. Thus we get cr(F)=ncr(F)=n, as desired.

Now consider the case n4n\geq 4 and FF is on the line 1\ell_{1}. Suppose that cr(F)=ncr(F)=n. This means there is a zero-dimensional subscheme Zn1Z\subset\mathbb{P}^{n-1} of degree nn such that IZann(F)I_{Z}\subset\operatorname{ann}(F). Let JJ be the ideal generated by the quadrics in ann(F)\operatorname{ann}(F). Comparing Hilbert functions, we see that [IZ]2=[J]2=[ann(F)]2[I_{Z}]_{2}=[J]_{2}=[\operatorname{ann}(F)]_{2}. By Proposition 4.6(v), we get that β1,3(S/ann(F))=1\beta_{1,3}(S/\operatorname{ann}(F))=1 and hence ann(F)\operatorname{ann}(F) is generated by JJ together with one cubic. Thus by Proposition 4.3 the value of the Hilbert function of R/JR/J in degree 3 is n+1n+1. Since ZZ has degree nn, its Hilbert function must be 1,n,n,1,n,n,\ldots. It follows that the Castelnuovo-Mumford regularity of IZI_{Z} is two. Hence IZI_{Z} is generated by quadrics, as is JJ. This implies IZ=JI_{Z}=J, which is a contradiction because the Hilbert function of R/IZR/I_{Z} in degree 3 is nn. This proves cr(F)n+1cr(F)\geq n+1. We get equality because wr(F)=n+1wr(F)=n+1 by Proposition 4.5.

Second, assume FF is a point on the line 2\ell_{2}, but not 𝒫\mathcal{P}. Since 2\ell_{2} meets 𝒞\mathcal{C} only in 𝒫\mathcal{P}, the ideal ann(F)\operatorname{ann}(F) is generated by quadrics (see Proposition 4.6). Hence Proposition 4.7 gives cr(F)n+1cr(F)\geq n+1, as above. In the proof of Proposition 4.6(vii), we constructed a Gorenstein scheme XX of degree n+1n+1 supported at one point with the property that IXann(F)I_{X}\subset\operatorname{ann}(F). Hence, by definition of the cactus rank, we get cr(F)degX=n+1cr(F)\leq\deg X=n+1. Together with the lower bound this yields cr(F)=n+1cr(F)=n+1. ∎

For the cactus rank of general cubics, the reader may look at [BR]. Now we show that the cactus rank and the Waring rank of symmetric cubics are usually equal, but that the difference in some cases can be made arbitrary large when increasing the number of variables.

Corollary 4.9.

The cactus rank and the Waring rank of any symmetric cubic FF in n3n\geq 3 variables are equal unless F=a0p13+a1np1p2F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2} with a10a_{1}\neq 0 and n4n\geq 4. In this case we have cr(F)=n+1<2(n1)=wr(F)cr(F)=n+1<2(n-1)=wr(F).

Proof.

Compare Propositions 4.5 and 4.8. ∎

Proposition 4.10.

For any symmetric cubic form F=a0p13+a1np1p2+a2n2p3F=a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3}, S/ann(F)S/\operatorname{ann}(F) satisfies the SLP.

Proof.

If we are at the cusp 𝒫=(1:0:0)\mathcal{P}=(1:0:0) of 𝒞\mathcal{C}, F=p13F=p_{1}^{3}, wr(F)=1wr(F)=1, the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,1,1,1)(1,1,1,1), [ann(F)]1=xixj,ij1[\operatorname{ann}(F)]_{1}=\langle x_{i}-x_{j},i\neq j\rangle_{1} and clearly S/ann(F)S/\operatorname{ann}(F) has the SLP.

If we are at the flex 𝒬=(2:3:1)\mathcal{Q}=(2:-3:1) of 𝒞\mathcal{C} i.e. F=2p133np1p2+n2p3F=2p_{1}^{3}-3np_{1}p_{2}+n^{2}p_{3} and n=3n=3 we have wr(F)=2wr(F)=2, the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,2,2,1)(1,2,2,1) (Proposition 4.3) and ann(F)\operatorname{ann}(F) is a complete intersection artinian ideal of codimension 2. Hence, it has the SLP (see [HMNW, Proposition 4.4]). If we are at the flex 𝒬\mathcal{Q} of 𝒞\mathcal{C} and n>3n>3, we have wr(F)=nwr(F)=n (Proposition 4.5), the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,n1,n1,1)(1,n-1,n-1,1) (Proposition 4.3), we easily check that {xix1}i=2,,n\{x_{i}-x_{1}\}_{i=2,\cdots,n} is a basis of S/ann(F)S/\operatorname{ann}(F) and x1x_{1} is an SL element since x132i:[S/ann(F)]i[S/ann(F)]3ix_{1}^{3-2i}:[S/\operatorname{ann}(F)]_{i}\longrightarrow[S/\operatorname{ann}(F)]_{3-i} is an isomorphism for i=0,1i=0,1.

Assume that (a0,a1,a2)(a_{0},a_{1},a_{2}) is neither the flex nor the cusp of 𝒞\mathcal{C}. According to Proposition 4.3 the Hilbert function of S/ann(F)S/\operatorname{ann}(F) is (1,n,n,1)(1,n,n,1) and we distinguish two cases:

Case 1: wr(F)=nwr(F)=n. In this case, [ann(F)]2=[IX]2[\operatorname{ann}(F)]_{2}=[I_{X}]_{2} where XX is the support of the Waring decomposition of FF. Therefore, any non-zero divisor on XX is an SL element.

Case 2: wr(F)>nwr(F)>n. We will show that 3 linear forms are sufficient to provide an SL elements for all of them. To prove that \ell is an SL element for S/ann(F)S/\operatorname{ann}(F) it suffices to show that :[S/ann(F)]1[S/ann(F)]2\ell:[S/\operatorname{ann}(F)]_{1}\longrightarrow[S/\operatorname{ann}(F)]_{2} is injective. The map :[S/ann(F)]1[S/ann(F)]2\ell:[S/\operatorname{ann}(F)]_{1}\longrightarrow[S/\operatorname{ann}(F)]_{2} fails injectivity if there is a linear form gSg\in S such that g=0\ell g=0 in [S/ann(F)]2[S/\operatorname{ann}(F)]_{2}, i.e. there is a linear form gSg\in S such that gF=0\ell g\circ F=0. So, our strategy is to study when the quadratic form F\ell\circ F drops rank. We first observe that (a0,a1,a2)𝒞(a_{0},a_{1},a_{2})\notin\mathcal{C}. We consider =i=1nxi\ell=\sum_{i=1}^{n}x_{i} and we will check when \ell is an SL element for S/ann(F)S/\operatorname{ann}(F). We claim that quadratic form

(a0p13+a1np1p2+a2n2p3)=3na0p12+a1n2p2+2na1p12+3a2n2p2=(3a0+2a1)np12+(a1+3a2)n2p2\begin{array}[]{rcl}\ell\circ(a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3})&=&3na_{0}p_{1}^{2}+a_{1}n^{2}p_{2}+2na_{1}p_{1}^{2}+3a_{2}n^{2}p_{2}\\ &=&(3a_{0}+2a_{1})np_{1}^{2}+(a_{1}+3a_{2})n^{2}p_{2}\end{array}

drops rank exactly when (a0,a1,a2)13(a_{0},a_{1},a_{2})\in\ell_{1}\cup\ell_{3}. Indeed, the matrix

Mq=[(a1+3a2)n2+(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(a1+3a2)n2+(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(a1+3a2)n2+(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(3a0+2a1)n(a1+3a2)n2+(3a0+2a1)n]M_{q}=\left[\begin{smallmatrix}(a_{1}+3a_{2})n^{2}+(3a_{0}+2a_{1})n&(3a_{0}+2a_{1})n&(3a_{0}+2a_{1})n&\cdots&(3a_{0}+2a_{1})n\\ (3a_{0}+2a_{1})n&(a_{1}+3a_{2})n^{2}+(3a_{0}+2a_{1})n&(3a_{0}+2a_{1})n&\cdots&(3a_{0}+2a_{1})n\\ (3a_{0}+2a_{1})n&(3a_{0}+2a_{1})n&(a_{1}+3a_{2})n^{2}+(3a_{0}+2a_{1})n&\cdots&(3a_{0}+2a_{1})n\\ \vdots&\vdots&\vdots&\vdots&\vdots\\ (3a_{0}+2a_{1})n&(3a_{0}+2a_{1})n&(3a_{0}+2a_{1})n&\cdots&(a_{1}+3a_{2})n^{2}+(3a_{0}+2a_{1})n\end{smallmatrix}\right]

associated to the quadratic form q=(3a0+2a1)np12+(a1+3a2)n2p2q=(3a_{0}+2a_{1})np_{1}^{2}+(a_{1}+3a_{2})n^{2}p_{2} has determinant

det(Mq)=3n2n(a1+3a2)n1(a0+a1+a2).\det(M_{q})=3n^{2n}(a_{1}+3a_{2})^{n-1}(a_{0}+a_{1}+a_{2}).

Therefore, the quadratic form (a0p13+a1np1p2+a2n2p3)\ell\circ(a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3}) drops rank exactly when (a0,a1,a2)13(a_{0},a_{1},a_{2})\in\ell_{1}\cup\ell_{3}, and we conclude that =i=1nxi\ell=\sum_{i=1}^{n}x_{i} is an SL element except on the lines 1\ell_{1} and 3\ell_{3}. Finally, we consider the quadratic forms

q1:=x1(a0p13+a1np1p2+a2n2p3)=3a0p12+2a1nx1p1+3a2n2x12+a1np2q_{1}:=x_{1}\circ(a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3})=3a_{0}p_{1}^{2}+2a_{1}nx_{1}p_{1}+3a_{2}n^{2}x_{1}^{2}+a_{1}np_{2}

and

q2:=(nx1p1)(a0p13+a1np1p2+a2n2p3)=2a1np12+2a1n2x1p1+3a2n3x123a2n2p2.q_{2}:=(nx_{1}-p_{1})\circ(a_{0}p_{1}^{3}+a_{1}np_{1}p_{2}+a_{2}n^{2}p_{3})=-2a_{1}np_{1}^{2}+2a_{1}n^{2}x_{1}p_{1}+3a_{2}n^{3}x_{1}^{2}-3a_{2}n^{2}p_{2}.

Computing the determinant of the symmetric matrices associated to q1q_{1} and q2q_{2} we get that the quadratic form q1q_{1} and q2q_{2} never drop rank simultaneously in points (a0,a1,a2)(a_{0},a_{1},a_{2}) of the lines 3\ell_{3} and 1\ell_{1}. Therefore on these lines either =x1\ell=x_{1} or =nx1ixi\ell=nx_{1}-\sum_{i}x_{i} is an SL element. ∎

5. Generic Waring rank for symmetric forms

As we have seen in the previous section, the generic symmetric cubic form in nn variables has Waring rank n+1n+1. We will now see that we can get bounds for the Waring rank of generic symmetric quartics and quintics using the same kind of orbits that come into the decompositions of Theorem 3.2. We start by looking at the power sum expansion

F(α0,α1,α2)=α03h13+i=1n(α1h1+α2Xi)3=α23p3+3α1α22p1p2+(α03+nα13+3α12α2)p13F(\alpha_{0},\alpha_{1},\alpha_{2})=\alpha_{0}^{3}h_{1}^{3}+\sum_{i=1}^{n}(\alpha_{1}h_{1}+\alpha_{2}X_{i})^{3}=\alpha_{2}^{3}p_{3}+3\alpha_{1}\alpha_{2}^{2}p_{1}p_{2}+(\alpha_{0}^{3}+n\alpha_{1}^{3}+3\alpha_{1}^{2}\alpha_{2})p_{1}^{3}

which gives rise to a rational map from 2\mathbb{P}^{2} with coordinates (α0:α1:α2)(\alpha_{0}\colon\alpha_{1}\colon\alpha_{2}) to 2(p3,p2p1,p13)\mathbb{P}_{2}(\langle p_{3},p_{2}p_{1},p_{1}^{3}\rangle). In order to see that this map is generically onto, we look at the Jacobian

Fiαj=[003α0203α223nα12+6α1α23α226α1α23α12]\frac{\partial F_{i}}{\partial\alpha_{j}}=\begin{bmatrix}0&0&3\alpha_{0}^{2}\\ 0&3\alpha_{2}^{2}&3n\alpha_{1}^{2}+6\alpha_{1}\alpha_{2}\\ 3\alpha_{2}^{2}&6\alpha_{1}\alpha_{2}&3\alpha_{1}^{2}\\ \end{bmatrix}

with determinant equal to 27α02α2427\alpha_{0}^{2}\alpha_{2}^{4}. This is non-zero except when α0=0\alpha_{0}=0, which corresponds to the cubic curve 𝒞\mathcal{C} and when α2=0\alpha_{2}=0, which corresponds to the point 𝒫\mathcal{P}, i.e., the cusp of 𝒞\mathcal{C}.

Proposition 5.1.

The generic symmetric quartic has Waring rank at most (n+12)+1=(n+22)n\binom{n+1}{2}+1=\binom{n+2}{2}-n and the generic symmetric quintic has Waring rank at most (n+22)\binom{n+2}{2}.

Proof.

We start by using the decomposition

F(α0,α1,α2,α3,α4)=α04h14+i=1n(α1h1+α2Xi)4+1i<jn(α3h1+α4Xi+α4Xj)4,F(\alpha_{0},\alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4})=\alpha_{0}^{4}h_{1}^{4}+\sum_{i=1}^{n}(\alpha_{1}h_{1}+\alpha_{2}X_{i})^{4}+\sum_{1\leq i<j\leq n}(\alpha_{3}h_{1}+\alpha_{4}X_{i}+\alpha_{4}X_{j})^{4},

which produces a symmetric quartic form of Waring rank at most 1+n+(n2)=(n+12)+1=(n+22)n1+n+\binom{n}{2}=\binom{n+1}{2}+1=\binom{n+2}{2}-n. We expand this into the basis of symmetric quartics given by {p4,p3p1,p22,p2p12,p14}\{p_{4},p_{3}p_{1},p_{2}^{2},p_{2}p_{1}^{2},p_{1}^{4}\} to get

F(α0,α1,α2,α3,α4)=(α24+(n8)α44)p4+(4α1α23+4(n4)α3α43+4α44)p3p1+3α44p22+(6α12α22+6(n2)α32α42+12α3α43)p2p12+(α04+nα14+4α13α2+(n2)α34+4(n1)α33α4+6α32α42)p14.\begin{split}F(\alpha_{0},\alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4})=(\alpha_{2}^{4}+(n-8)\alpha_{4}^{4})p_{4}+(4\alpha_{1}\alpha_{2}^{3}+4(n-4)\alpha_{3}\alpha_{4}^{3}+4\alpha_{4}^{4})p_{3}p_{1}+3\alpha_{4}^{4}p_{2}^{2}\\ +(6\alpha_{1}^{2}\alpha_{2}^{2}+6(n-2)\alpha_{3}^{2}\alpha_{4}^{2}+12\alpha_{3}\alpha_{4}^{3})p_{2}p_{1}^{2}\\ +(\alpha_{0}^{4}+n\alpha_{1}^{4}+4\alpha_{1}^{3}\alpha_{2}+\binom{n}{2}\alpha_{3}^{4}+4(n-1)\alpha_{3}^{3}\alpha_{4}+6\alpha_{3}^{2}\alpha_{4}^{2})p_{1}^{4}.\end{split}

Thus we have produced a rational map from 4\mathbb{P}^{4} with coordinates (α0:α1:α2:α3:α4)(\alpha_{0}\colon\alpha_{1}\colon\alpha_{2}\colon\alpha_{3}\colon\alpha_{4}) to 4=(p4,p3p1,p22,p2p12,p14)\mathbb{P}^{4}=\mathbb{P}(\langle p_{4},p_{3}p_{1},p_{2}^{2},p_{2}p_{1}^{2},p_{1}^{4}\rangle). In order to show that this is generically onto, which proves our statement, we consider the Jacobian

(Fjαi)=[00004α0304α23012α1α224nα13+12α12α24α2312α1α22012α12α24α1304(n4)α43012(n2)α3α42+12α432n(n1)α33+12(n1)α32α4+12α3α424(n8)α4312(n4)α3α42+16α4312α4312(n2)α32α4+36α3α424(n1)α33+12α32α4]\left(\frac{\partial F_{j}}{\partial\alpha_{i}}\right)=\left[\begin{smallmatrix}0&0&0&0&4\alpha_{0}^{3}\\ 0&4\alpha_{2}^{3}&0&12\alpha_{1}\alpha_{2}^{2}&4n\alpha_{1}^{3}+12\alpha_{1}^{2}\alpha_{2}\\ 4\alpha_{2}^{3}&12\alpha_{1}\alpha_{2}^{2}&0&12\alpha_{1}^{2}\alpha_{2}&4\alpha_{1}^{3}\\ 0&4(n-4)\alpha_{4}^{3}&0&12(n-2)\alpha_{3}\alpha_{4}^{2}+12\alpha_{4}^{3}&2n(n-1)\alpha_{3}^{3}+12(n-1)\alpha_{3}^{2}\alpha_{4}+12\alpha_{3}\alpha_{4}^{2}\\ 4(n-8)\alpha_{4}^{3}&12(n-4)\alpha_{3}\alpha_{4}^{2}+16\alpha_{4}^{3}&12\alpha_{4}^{3}&12(n-2)\alpha_{3}^{2}\alpha_{4}+36\alpha_{3}\alpha_{4}^{2}&4(n-1)\alpha_{3}^{3}+12\alpha_{3}^{2}\alpha_{4}\end{smallmatrix}\right]

which has determinant

det(Fjαi)=(4α03)(4α23)(12α43)det[4α2312α1α224(n4)α4312(n2)α3α42+12α43]=21032α03α25α45((n2)α2α3(n4)α1α4+α2α4).\begin{split}\det\left(\frac{\partial F_{j}}{\partial\alpha_{i}}\right)=(4\alpha_{0}^{3})(4\alpha_{2}^{3})(12\alpha_{4}^{3})\det\left[\begin{matrix}4\alpha_{2}^{3}&12\alpha_{1}\alpha_{2}^{2}\\ 4(n-4)\alpha_{4}^{3}&12(n-2)\alpha_{3}\alpha_{4}^{2}+12\alpha_{4}^{3}\end{matrix}\right]\\ =2^{10}3^{2}\alpha_{0}^{3}\alpha_{2}^{5}\alpha_{4}^{5}((n-2)\alpha_{2}\alpha_{3}-(n-4)\alpha_{1}\alpha_{4}+\alpha_{2}\alpha_{4}).\end{split}

This is generically non-zero, showing that the generic symmetric quartic has power sum expansion with at most (n+12)+1\binom{n+1}{2}+1 terms.

For the symmetric quintics, we use the power sum expansion

F(α0,α1,α2,α3,α4,α5,α6)=α05h15+i=1n(α1h1+α2Xi)5+1i<jn(α3h1+α4Xi+α4Xj)5+i=1n(α5h1+α6Xi)5\begin{split}F(\alpha_{0},\alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4},\alpha_{5},\alpha_{6})=\alpha_{0}^{5}h_{1}^{5}+\sum_{i=1}^{n}(\alpha_{1}h_{1}+\alpha_{2}X_{i})^{5}+\sum_{1\leq i<j\leq n}(\alpha_{3}h_{1}+\alpha_{4}X_{i}+\alpha_{4}X_{j})^{5}\\ +\sum_{i=1}^{n}(\alpha_{5}h_{1}+\alpha_{6}X_{i})^{5}\end{split}

which gives a rational map from 6\mathbb{P}^{6} to 6=(p5,p4p1,p3p2,p3p12,p22p1,p2p13,p15)\mathbb{P}^{6}=\mathbb{P}(\langle p_{5},p_{4}p_{1},p_{3}p_{2},p_{3}p_{1}^{2},p_{2}^{2}p_{1},p_{2}p_{1}^{3},p_{1}^{5}\rangle) given by the expansion

F(α0,α1,,α6)=(α25+(16n)α45+α65)p5+(5α1α24+5(n8)α3α44+5α45+5α5α54)p4p1+10α45p3p2+(10α13α23+10(n4)α32α43+20α3α44+10α52α63)p3p12+15α3α44p22p1+(10α13α22+10(n2)α33α42+30α32α43+10α53α62)p2p13+(α05+nα15+5α14α2+5(n1)α34α4+10α33α42+5α54α6+(n2)α35)p15.\begin{split}F(\alpha_{0},\alpha_{1},\dots,\alpha_{6})=(\alpha_{2}^{5}+(16-n)\alpha_{4}^{5}+\alpha_{6}^{5})p_{5}+(5\alpha_{1}\alpha_{2}^{4}+5(n-8)\alpha_{3}\alpha_{4}^{4}+5\alpha_{4}^{5}+5\alpha_{5}\alpha_{5}^{4})p_{4}p_{1}\\ +10\alpha_{4}^{5}p_{3}p_{2}+(10\alpha_{1}^{3}\alpha_{2}^{3}+10(n-4)\alpha_{3}^{2}\alpha_{4}^{3}+20\alpha_{3}\alpha_{4}^{4}+10\alpha_{5}^{2}\alpha_{6}^{3})p_{3}p_{1}^{2}\\ +15\alpha_{3}\alpha_{4}^{4}p_{2}^{2}p_{1}+(10\alpha_{1}^{3}\alpha_{2}^{2}+10(n-2)\alpha_{3}^{3}\alpha_{4}^{2}+30\alpha_{3}^{2}\alpha_{4}^{3}+10\alpha_{5}^{3}\alpha_{6}^{2})p_{2}p_{1}^{3}\\ +(\alpha_{0}^{5}+n\alpha_{1}^{5}+5\alpha_{1}^{4}\alpha_{2}+5(n-1)\alpha_{3}^{4}\alpha_{4}+10\alpha_{3}^{3}\alpha_{4}^{2}+5\alpha_{5}^{4}\alpha_{6}+\binom{n}{2}\alpha_{3}^{5})p_{1}^{5}.\end{split}

The Jacobian of this map is given by

[0000005α0405α24020α1α23030α12α225nα14+20α13α25α2420α1α23030α12α22020α13α25α1405(n8)α44020(n4)α3α43+20α4415α4430(n2)α32α42+60α3α435(n2)α34+20(n1)α33α4+30α32α425(n16)α4420(n8)α3α43+25α4450α4430(n4)α32α42+80α3α4360α3α4320(n2)α33α4+90α32α425(n1)α34α4+20α33α405α64020α5α63030α52α625nα54+20α53α65α6420α5α63030α52α62020α53α65α54]\left[\begin{smallmatrix}0&0&0&0&0&0&5\alpha_{0}^{4}\\ 0&5\alpha_{2}^{4}&0&20\alpha_{1}\alpha_{2}^{3}&0&30\alpha_{1}^{2}\alpha_{2}^{2}&5n\alpha_{1}^{4}+20\alpha_{1}^{3}\alpha_{2}\\ 5\alpha_{2}^{4}&20\alpha_{1}\alpha_{2}^{3}&0&30\alpha_{1}^{2}\alpha_{2}^{2}&0&20\alpha_{1}^{3}\alpha_{2}&5\alpha_{1}^{4}\\ 0&5(n-8)\alpha_{4}^{4}&0&20(n-4)\alpha_{3}\alpha_{4}^{3}+20\alpha_{4}^{4}&15\alpha_{4}^{4}&30(n-2)\alpha_{3}^{2}\alpha_{4}^{2}+60\alpha_{3}\alpha_{4}^{3}&5\binom{n}{2}\alpha_{3}^{4}+20(n-1)\alpha_{3}^{3}\alpha_{4}+30\alpha_{3}^{2}\alpha_{4}^{2}\\ 5(n-16)\alpha_{4}^{4}&20(n-8)\alpha_{3}\alpha_{4}^{3}+25\alpha_{4}^{4}&50\alpha_{4}^{4}&30(n-4)\alpha_{3}^{2}\alpha_{4}^{2}+80\alpha_{3}\alpha_{4}^{3}&60\alpha_{3}\alpha_{4}^{3}&20(n-2)\alpha_{3}^{3}\alpha_{4}+90\alpha_{3}^{2}\alpha_{4}^{2}&5(n-1)\alpha_{3}^{4}\alpha_{4}+20\alpha_{3}^{3}\alpha_{4}\\ 0&5\alpha_{6}^{4}&0&20\alpha_{5}\alpha_{6}^{3}&0&30\alpha_{5}^{2}\alpha_{6}^{2}&5n\alpha_{5}^{4}+20\alpha_{5}^{3}\alpha_{6}\\ 5\alpha_{6}^{4}&20\alpha_{5}\alpha_{6}^{3}&0&30\alpha_{5}^{2}\alpha_{6}^{2}&0&20\alpha_{5}^{3}\alpha_{6}&5\alpha_{5}^{4}\\ \end{smallmatrix}\right]

and we get the determinant of that as

(5α04)(15α44)(50α4)4det[05α2420α1α2330α12α225α2420α1α2330α12α2220α13α205α6420α5α6330α52α625α6420α5α6330α52α6220α53α6]=23359α04α24α48α64(α1α6α2α5)4.(5\alpha_{0}^{4})(15\alpha_{4}^{4})(50\alpha_{4})^{4}\det\left[\begin{smallmatrix}0&5\alpha_{2}^{4}&20\alpha_{1}\alpha_{2}^{3}&30\alpha_{1}^{2}\alpha_{2}^{2}\\ 5\alpha_{2}^{4}&20\alpha_{1}\alpha_{2}^{3}&30\alpha_{1}^{2}\alpha_{2}^{2}&20\alpha_{1}^{3}\alpha_{2}\\ 0&5\alpha_{6}^{4}&20\alpha_{5}\alpha_{6}^{3}&30\alpha_{5}^{2}\alpha_{6}^{2}\\ 5\alpha_{6}^{4}&20\alpha_{5}\alpha_{6}^{3}&30\alpha_{5}^{2}\alpha_{6}^{2}&20\alpha_{5}^{3}\alpha_{6}\\ \end{smallmatrix}\right]=2^{3}\cdot 3\cdot 5^{9}\alpha_{0}^{4}\alpha_{2}^{4}\alpha_{4}^{8}\alpha_{6}^{4}(\alpha_{1}\alpha_{6}-\alpha_{2}\alpha_{5})^{4}.

This is generically non-zero. In fact, it is non-zero as long as the linear forms involved in the decomposition are distinct. Thus the generic symmetric quintic form has Waring rank at most (n+22)\binom{n+2}{2}. ∎

Remark 5.2.

In the case of cubics, Proposition 4.5 shows that the generic rank is indeed n+1n+1. In the case of quartics, we know from the Hilbert function that the generic rank is at least (n+12)\binom{n+1}{2} and our upper bound is just one more. However, we are not able to use this approach to prove that the generic rank of symmetric quartics cannot be (n+12)\binom{n+1}{2}. Observe that the generic symmetric form might have a Waring decomposition that is not invariant under the action of the symmetric group. In the case of quintics, the lower bound given by the Hilbert function is again (n+12)\binom{n+1}{2} and our upper bound is n+1n+1 higher than this. Here we might be able to use the resolution in order to show that the generic rank is higher than (n+12)\binom{n+1}{2}, but we have not been able to do this in general.

For symmetric forms of degree six and higher, we cannot expect that the generic symmetric form can be expanded into powers of linear forms in a similar way as the complete symmetric form. This can be seen from looking at the dimension of the family of symmetric forms compared to the number of parameters that can be involved in the expansions. For example, in the case of sextics, the dimension of the family is ten while an expansion corresponding to the expansion we have for quartics would give only a nine-dimensional family. For higher degrees, the difference in dimensions grows larger and larger.

Remark 5.3.

We can now look more closely at the claims of Remark 3.3. Using the rational map from 4\mathbb{P}^{4} to 4\mathbb{P}^{4} described in the proof of Proposition 5.1 we can look at the preimage of hn,4h_{n,4}. Using the fact that

hn,4=124(6p4+8p3p1+3p22+6p2p12+p14)h_{n,4}=\frac{1}{24}(6p_{4}+8p_{3}p_{1}+3p_{2}^{2}+6p_{2}p_{1}^{2}+p_{1}^{4})

we get that the preimage is given by the following system of equations

{α24+(n14)α44=0α1α23α44+(n4)α3α43=0α12α22+2α3α43+(n2)α32α42α44=0α04+4α13α2+nα14+6α32α42+4(n1)α33α4+n(n1)/2α34α44=0.\left\{\begin{array}[]{rcl}\alpha_{2}^{4}+(n-14)\alpha_{4}^{4}&=&0\\ \alpha_{1}\alpha_{2}^{3}-\alpha_{4}^{4}+(n-4)\alpha_{3}\alpha_{4}^{3}&=&0\\ \alpha_{1}^{2}\alpha_{2}^{2}+2\alpha_{3}\alpha_{4}^{3}+(n-2)\alpha_{3}^{2}\alpha_{4}^{2}-\alpha_{4}^{4}&=&0\\ \alpha_{0}^{4}+4\alpha_{1}^{3}\alpha_{2}+n\alpha_{1}^{4}+6\alpha_{3}^{2}\alpha_{4}^{2}+4(n-1)\alpha_{3}^{3}\alpha_{4}+n(n-1)/2\alpha_{3}^{4}-\alpha_{4}^{4}&=&0.\\ \end{array}\right.

From the first three equations, we get

α12α26=(α44(n4)α3α43)2=(14n)α44(α442α3α43(n2)α32α42)\alpha_{1}^{2}\alpha_{2}^{6}=(\alpha_{4}^{4}-(n-4)\alpha_{3}\alpha_{4}^{3})^{2}=(14-n)\alpha_{4}^{4}(\alpha_{4}^{4}-2\alpha_{3}\alpha_{4}^{3}-(n-2)\alpha_{3}^{2}\alpha_{4}^{2})

which gives

α46(α42n4α4α3n+8α32n13α42+36α4α312α32)=0.\alpha_{4}^{6}(\alpha_{4}^{2}n-4\alpha_{4}\alpha_{3}n+8\alpha_{3}^{2}n-13\alpha_{4}^{2}+36\alpha_{4}\alpha_{3}-12\alpha_{3}^{2})=0.

For all nn we can solve this equation with α40\alpha_{4}\neq 0 and hence get solutions for α0\alpha_{0}, α1\alpha_{1} and α2\alpha_{2} in terms of α4\alpha_{4}. Because of the first equation, we see that α20\alpha_{2}\neq 0 when n14n\neq 14. When n=14n=14, we get that α2=0\alpha_{2}=0 from the first equation and then we can see that the Waring rank of F(α0,α1,0,α3,α4)F(\alpha_{0},\alpha_{1},0,\alpha_{3},\alpha_{4}) is less than the Waring rank of h14,4h_{14,4} showing that the solution to our equations give F(α0,α1,0,α3,α4)=0F(\alpha_{0},\alpha_{1},0,\alpha_{3},\alpha_{4})=0.

Generic rank in the case n=14n=14 is 170170 which is far larger than 120120 given by the decomposition in our theorem. However, it is unclear if we can find Waring decompositions with fewer terms. If we look for symmetric Waring decompositions, we need to take a union of orbits and it will not be sufficient to take nn-point orbits since such decompositions do not give any contribution to the coefficient of p22p_{2}^{2}.

For n=6n=6, the generic rank is lower than the bound given by our construction, 2121 instead of 2222. Also for n<6n<6, generic rank is lower by one. For n>6n>6, generic rank is always higher.

References

  • [AH] J. Alexander and A. Hirschowitz, Polynomial interpolation in several variables, J. Algebraic Geom. 4 (1995), no. 2, 201–222.
  • [BCGM] M. Beltrametti, E. Carletti, D. Gallarati and G. Monti Bragadin, Lectures on curves, surfaces and projective varieties. A classical view of algebraic geometry, EMS Textbooks in Mathematics. European Mathematical Society (EMS), Zürich, 2009.
  • [BR] A. Bernardi and K. Ranestad, On the cactus rank of cubic forms, Journal of Symbolic Computation 50 (2013) 291–297.
  • [B] M. Boij, Gorenstein Artin algebras and points in projective space, Bull. London Math. Soc. 31 (1999) 11–16.
  • [FL] R. Fröber and D. Laksov, Compressed algebras. Complete intersections (Acireale, 1983), 121–151 Lecture Notes in Math. 1092, Springer, Berlin, 1984.
  • [HMNW] T. Harima, J. Migliore, U. Nagel, and J. Watanabe, The Weak and Strong Lefschetz properties for Artinian K-algebras, J. Algebra 262 (2003), 99–126.
  • [HK] J. Herzog and M. Kühl, On the Betti numbers of finite pure and linear resolutions, Comm. Algebra 12 (1984), no. 13-14, 1627–-1646.
  • [I] A. Iarrobino, Inverse system of a symbolic power. II. The Waring problem for forms, J. Algebra, 174 (1995), no. 3, 1091–1110.
  • [IK] A. Iarrobino and V. Kanev, Power sums, Gorenstein algebras, and determinantal loci, Lecture Notes in Math. 1721, Springer, Berlin, 1999.
  • [GL] F. Gesmundo and J. M. Landsberg, Explicit polynomial sequences with maximal spaces of partial derivates and a question of K. Mulmuley, Theory Comput. 15 (2019), Paper No. 3, 24 pp.
  • [RS] K. Ranestad and F.-O. Schreyer, On the rank of a symmetric form, J. Algebra 346 (2011), 340–342.