Waring and cactus ranks and Strong Lefschetz Property for annihilators of symmetric forms
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.
1. Introduction
Let be a field of characteristic and consider the polynomial ring and its inverse system ; more precisely, is an -module where acts as . It is a well-known fact ([IK, Lemma 3.12]) that if is a general polynomial of degree in then the annihilator in defines an artinian Gorenstein algebra with compressed Hilbert function, meaning that the “first half” of the Hilbert function of agrees with the Hilbert function of , and the second half is of course determined by symmetry.
This begs the question of how “general” 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 ) 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 is the complete symmetric polynomial then has compressed Hilbert function. (The fact that 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 even satisfies the Strong Lefschetz Property (SLP). We recall that a graded artinian -algebra satisfies the Weak Lefschetz Property (WLP) if multiplication by a general linear form from any component to the next has maximal rank, and it satisfies SLP if maximal rank also holds for for all and . In the same section we also give a useful decomposition of (Theorem 2.12), separating into the cases where 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 is a homogeneous polynomial of degree , recall that its Waring rank is the smallest so that can be written in the form for linear forms , and a power sum decomposition is any expression of in this form (not necessarily minimal). Let us denote by the complete symmetric polynomial of degree in variables. The main result of Section 3 is a specific power sum decomposition for and separately for , for any and . In both the even and odd cases, our power sum decomposition has terms, giving an upper bound for the Waring rank. Thus our bound is a polynomial of degree in . Note that the Waring rank for general forms of degree grows like a polynomial of degree in (see Remark 3.3). Later in the paper we will show that our bound for 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, . Inside of this plane, we consider the forms that have Waring rank at most . We show that this locus can be viewed as a rational cuspidal cubic curve, , and we give its equation (Lemma 4.2). From our perspective there are two important points on : the (unique) flex point and the cusp . Then there are three important lines: the line joining and , the tangent line to at (meaning the unique line that meets at and no other point), and the tangent to at . We give the equations of these three lines.
With this preparation, we divide the points of as follows: , , the remaining points of , the remaining points of , the remaining points of , the remaining points of , and the remaining points of . For each point of , depending on which of these categories it falls into, we give the Hilbert function of the corresponding algebra, a description of the generators of , 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 is the smallest degree of a zero-dimensional subscheme whose saturated ideal is contained in . We show as a consequence that for symmetric cubic forms, the Waring rank and the cactus rank are equal except for points on other than (Corollary 4.9). We end this section by showing that for any symmetric cubic form , the corresponding algebra 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 be a field of characteristic and consider the polynomial ring and its inverse system which means that is an -module where acts as .
The homogeneous, or complete, symmetric polynomials are defined as
The goal of this section is to prove that the Gorenstein algebra is compressed and satisfies the SLP (Theorem 2.11). The fact that it is compressed was also shown in [GL].
We fix and in the formal power series ring , which is also a graded -module, we define . We will often use to denote the linear form in .
Lemma 2.1.
The inverse system of the ideal is given by the subring considered as an -module.
Proof.
The linear form annihilates any polynomial in and the dimension of the inverse system is in degree which equals the dimension of . Thus is the inverse system of the ideal . ∎
Lemma 2.2.
Proof.
The monomial of degree occurs from the derivatives . The sum of these contributions is . ∎
Definition 2.3.
Let be the -linear map defined on the monomial basis by
for each .
Remark 2.4.
Note that the inverse image of the monomial basis of gives the dual basis with respect to the action. Moreover, powers of are sent to multiples of the homogeneous symmetric polynomials, .
Definition 2.5.
For any we define
Lemma 2.6.
For any we have that
Proof.
We expand by the binomial theorem as
Now look at the contribution to by each term of this sum. The only monomials in that contribute are the multiples of the corresponding monomial in the variables . We get that
Summing over the terms of now gives
which concludes the proof of the lemma. ∎
Lemma 2.7.
For in with we have that
-
(1)
, for .
-
(2)
.
-
(3)
, where
Proof.
The first two statements follow from the fact that the initial term in is which has degree . The last statement is given by the power series expansion of
∎
Lemma 2.8.
For and in with , we have that
Proof.
We have
where we use that
which comes from looking at the coefficient of in . ∎
Definition 2.9.
For integers let
Proposition 2.10.
The pairing given by is trivial when and is perfect when .
Proof.
We have that a basis for is given by with . We have that
If , we have that either or annihilates by Lemma 2.7(1) since either or . Hence the pairing is trivial.
If , it suffices to show that the pairing given by is perfect. By Lemma 2.7 (2) we get that for in we have . Hence the pairing is perfect since is invertible. ∎
Theorem 2.11.
For each the Gorenstein algebra is compressed and satisfies the SLP.
Proof.
For even socle degree, , we need to show that the ideal is zero in degree in order to conclude that the algebra is compressed. In degree , we have that . Using Proposition 2.10 we see that the pairing given by 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 , we conclude that the pairing is perfect on , which implies that is trivial in degree .
For odd socle degree, , it is enough to show that is zero in degree . If has degree and satisfies we also have by Lemma 2.2, but we have shown that is trivial in degree and therefore . Hence the algebra is compressed.
Let . In order to prove that has the strong Lefschetz property, it is sufficient to show that is an isomorphism for all . Assume that satisfies in . This means that , which by Lemma 2.2 means that . However, since is compressed, this means that and the multiplication by is injective and by symmetry of the Hilbert function also bijective. ∎
Theorem 2.12.
For even socle degree, , we have that and for odd socle degree, , we have that .
Proof.
Since the algebra is compressed of even socle degree we have generators of only in degree and the number of generators is given by
Both and annihilate and their intersection is trivial as we can see using the pairing given by . Indeed, from Proposition 2.10 we have that for a non-zero element in , there is an element with , while if , we have to have from the same proposition.
Moreover, and , which proves that .
Since the algebra is compressed of odd socle degree and we can have generators in degree and . The dimension of in degree is given by which equals . Since annihilates we must have that in degree equals . Let be an element of . Since we have that we must have that for some and in . Now implies that . From above it follows that so . Thus we see that , 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 we have that
Proof.
Assume that and , where . Then or , where and . Applying we get
and division by shows that . Hence . The other direction is trivial. ∎
In the following we will use the notation for the complete symmetric polynomial of degree in the variables .
Theorem 3.2.
For and the following power sum decompositions hold for the complete symmetric functions
and
where in both cases the second summation contains just one term and , respectively, for .
Proof.
We will prove this by induction on and and we introduce the notation
as an expression valid in any polynomial ring with . Furthermore, for , we introduce the notation
for , and . We shall prove by induction also on that
(1) |
The statement of the theorem is (1) in the case .
The base of the induction is given by the cases , and . For we have that and we get
so the assertions hold for . For we have that and we get
and
according to the previous equality. Thus both equalities hold for .
For we have that
and the coefficient of equals the coefficient of in the formal power series
which is equal to
For we get
which is the coefficient of in the formal power series
which equals
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 and after restriction to .
We have that
where we have to assume by induction that the equality holds for . We have that
We subtract , which by induction on is zero, from this and get
which agrees with the derivative of the right-hand side of (1).
Next, we restrict to where we have . Hence
and
∎
Remark 3.3.
For either even degree or odd degree , our power sum decomposition in Theorem 3.2 has 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 have Waring rank and general forms of degree have Waring rank . Observe that our bound has degree as a polynomial in while the Waring rank for general forms grows like a polynomial of degree or in . 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 (i.e. degree 3) in the next section. For even degree, experimentally we have seen that it is not necessarily sharp.
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 states that
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 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 , where are the power sum symmetric forms in . Let 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 . For in , define (up to a scalar multiple) a symmetric cubic
(2) | ||||
Furthermore, let be the rational cubic curve defined as the image of the map
(Note the coefficients and on the basis.) That means, the image of corresponds to the cubic .
We also will use the above coordinates to parameterize any symmetric cubic in variables. In fact, since the power sums whose degrees are at most generate the ring of symmetric polynomials, every symmetric cubic can be written (over a field of characteristic zero) as . We refer to as the cubic at the point .
Notation 4.1.
We fix three lines in which will play an important role in this section:
is the line defined by ,
is the line defined by , and
is the line defined by .
Lemma 4.2.
The cubic curve is a cuspidal curve given by the equation
Its cusp is and its unique flex point is . The line through and is the line . Moreover, is the unique line meeting at with multiplicity three.
Proof.
It is elementary to check that has a singularity at . This is a cusp rather than a node because there is only one point in such that , namely . 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 is the line and that the intersection of this line with at has multiplicity 3. Hence is indeed a flex point. The rest is immediate. ∎
Below we slightly abuse notation and refer to the line as the tangent line to at .
Proposition 4.3.
The Hilbert function of is
Proof.
Since is -invariant there are three possibilities for , where is the trivial representation and is the standard representation. If , we have or . In the second case, and and in the first case . There is a unique solution to since
In all other points . ∎
In order to find the Waring rank of all symmetric cubic forms we need the following lemma about small orbits in under the action of . We will assume that .
Lemma 4.4.
If is an orbit of points in under the action of with , we have that one of the following holds
-
(1)
and .
-
(2)
and , .
-
(3)
and , where .
-
(4)
and .
Proof.
If , we have at least one point of the form . If not all are equal, we have at least points of this form. In the case , this gives a possible orbit if or . In the first case, and we get or . In the second case, we get , , so . Thus we are we are either in case (1) or case (3). When , we get a possible orbit if . If we must have or . In the first case we get and we are in case (2). In the second case we get and which gives case (4). If , we have to have and which leads to the case (2). ∎
Proposition 4.5.
The Waring rank of the symmetric cubic form at is
Proof.
The Waring rank is if and only if the Hilbert function is and we are at which is the cusp of according to Lemma 4.2.
We have that the Waring rank has to be at least at the flex point and at least at all other points. If the rank is at the flex point we would have where is the support of the Waring decomposition. Since is symmetric, has to be a union of orbits and Lemma 4.4 says that we can only have or . In the case , we do get a Waring decomposition of as
In the case , there is no non-zero symmetric cubic that can be expressed in terms of the three linear forms . Thus, for , the Waring rank has to be at the flex point.
For points on the curve we have a power sum decomposition with terms (see Equation (2)). Moreover, for all these points apart from the flex point and the cusp, the Hilbert function of is by Proposition 4.3. Hence the Waring rank of has to be equal to .
For points on the line , which is the tangent line to at the cusp, we have that . Let . We have that and as an affine variety. Let . If has a Waring decomposition corresponding to the reduced set of points , we have that . We hence have that
which gives
This shows that . We will now show that has a Waring decomposition of length . Let and let denote the ideal in generated by , for . The ideal in generated by the same forms is an artinian Gorenstein ideal given by the annihilator of . Thus is a zero-dimensional Gorenstein scheme of length concentrated at the point . Moreover, it is contained in . We now take a Waring decomposition of which has length since is a non-degenerate quadric in variables. This gives a reduced subscheme of length and the cone over this is a reduced curve consisting of lines meeting at the point . The ideal is artinian and contains . Hence Bertini’s theorem says that we can find a quadric in intersecting in distinct points. The ideal of these points is contained in which shows that .
It remains to show that away from and its tangent line at the cusp. Through any such point, we have a line through the cusp, meeting the curve in exactly one more point. This gives a decomposition into cubes of linear forms. Since the Hilbert function of is , and if it is equal to , the ideal of the corresponding points equals the ideal in degree . Hence the points are invariant under the action of and by Lemma 4.4 the only possible orbits give points on the curve . ∎

Before looking at the cactus rank, we will study the structure of the resolutions of for symmetric cubics.
Proposition 4.6.
For any symmetric cubic the Betti numbers of can be obtained from the Betti numbers of the coordinate ring of a zero-dimensional locally Gorenstein scheme . We can write , where is a canonical ideal and for and all we have that
There are the following cases for the resolution of .
-
For at , is a single point and the resolution of is linear with , for .
-
For at , and , is two points and the resolution of is a Koszul complex with .
-
For at , and , is an arithmetically Gorenstein set of points in a hyperplane and the resolution is forced by the -vector , i.e., and
-
For on }, and , is a set of three points not on a line, with a linear resolution, i.e., , and .
-
For on , and , is a set of points where of them are linearly general in a hyperplane. The resolution of is linear except for the last two homological degrees where . The Betti numbers are
-
For at away from and , is a set of points not in a hyperplane. The resolution of is linear with and
-
For on , is a non-degenerate arithmetically Gorenstein scheme of length concentrated at a single point and has an almost linear resolution, i.e., and
-
For all other , is an arithmetically Gorenstein set of points and has an almost linear resolution, i.e., and
Proof.
The existence of the set comes from the Waring decompositions of Proposition 4.5 except for in case and we will start by looking at what happens in that case. Here we have for some coefficients , where . Let be the ideal generated by the quadratic polynomials in the subring that annihilate . All such generators annihilate and , so for any . Furthermore, is artinian with Hilbert function . It is Gorenstein, since there is no socle in degree one. Indeed, which gives a non-degenerate pairing on . Thus we have that where is a zero-dimensional arithmetically Gorenstein subscheme concentrated at the point . In fact, is the subring of generated by . The symmetric quadric in is given by
and evaluation at gives
showing that is a non-zero-divisor on . Since is arithmetically Gorenstein, the ideal is a canonical ideal and we have .
In all the other cases, the Castelnuovo–Mumford regularity of the set of points is at most and we can apply [B, Theorem 3.4] to conclude that the image of the ideal in the coordinate ring of is a canonical ideal. When the Castelnuovo–Mumford regularity of 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 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 , is the quotient of the coordinate ring of by a non-zero-divisor and the ideal is still a canonical ideal since is arithmetically Gorenstein.
It remains to treat the case where the Castelnuovo–Mumford regularity of is two but the ideal is not given by a non-zero-divisor. We look at the short exact sequence
where is the canonical ideal that is the image of in . Since there may be cancellations, we only get inequalities
but in our case, the only possible cancellation would be in degree and dually in degree . If this cancellation occurred, it would mean that the ideal was generated by quadrics, but this is not possible, since the symmetric quadric is not a non-zero-divisor on since it vanishes at the point .
For each of the cases, we will now prove the statements about the resolutions of .
-
For , is the single point and the ideal is generated by the linear forms . Hence the resolution is linear and given by a Koszul complex of length , which gives the stated Betti numbers.
-
Here the Waring rank is two and we have that the ideal of two points in is a complete intersection of type . The resolution is a Koszul complex which gives the stated Betti numbers.
-
The Waring rank is and the set of points giving the Waring decomposition is the -orbit of . They lie in a hyperplane and are in linearly general position in that hyperplane. Hence they form an arithmetically Gorenstein set of points with -vector . The Betti numbers of the coordinate ring are given as sums of the consecutive Betti numbers of an artinian Gorenstein algebra with the same -vector. This means that and
where we get the Betti numbers for the artinian Gorenstein algebra with -vector by the formula for the Betti numbers of a pure resolution ([HK]).
-
We have that , where and the three points do not lie on a line. For three points not on a line in , the resolution is linear and the Betti numbers are , and .
-
is the union of the single point orbit and the -point orbit of . Since lies in a hyperplane, but is in linearly general position within that hyperplane, is arithmetically Gorenstein with -vector . Thus the -vector of has to be . For an artinian reduction of , the ideal in degree equals the ideal of an artinian reduction of . Thus we see that has a one-dimensional socle in degree and a one-dimensional socle in degree . This gives us the last column in the Betti table of and forces that . By duality, we get that for . Hence all the remaining Betti numbers are equal to the Betti numbers of an arithmetically Gorenstein scheme with the same -vector as , i.e., the same as in case below.
-
Here is the -orbit of where and . Thus is a set of points not in a hyperplane with -vector showing that the resolution is linear. Again, we can use the formula for the Betti numbers of a pure resolution to write
-
As we have seen, is an arithmetically Gorenstein subscheme of length concentrated at the point . The -vector is and the resolution of is almost linear since the artinian reduction is extremely compressed ([FL]). From the formula for the Betti numbers of a pure resolution, we get
-
Here is the union of the single point orbit and the -point orbit of where and . This set of points is in linearly general position with -vector . Thus is arithmetically Gorenstein and has an almost linear resolution as in the case 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 be a homogeneous polynomial. If the degree of any minimal generator of is at most , then one has
Proposition 4.8.
The cactus rank of the symmetric cubic form at is
Proof.
We will combine the lower bound provided by Proposition 4.7 and the upper bound , using that we know the Waring rank by Proposition 4.5. We consider cases depending on the Hilbert function of .
By Proposition 4.5, we know if and only if is the form at , i.e., , which has cactus rank one.
Now consider at the flex point of . If then is a complete intersection of type by Proposition 4.6(ii). Hence, this ideal does not contain the ideal of a point, which implies . Since by Proposition 4.5, we get . If , then is generated by quadrics (see Proposition 4.6(iii)) and has degree by Proposition 4.3. Hence Proposition 4.7 gives . We get equality because Proposition 4.5 yields .
It remains to consider cubics at points other than the cusp and the flex point . Hence, the Hilbert function of is by Proposition 4.3. In particular, has degree . Observe that is generated by quadrics, unless corresponds to a point on the curve or a point on the line . Excluding the latter two cases, Proposition 4.7 gives . We get equality if is also not on the line because then by Proposition 4.5. Thus, we are left with considering three situations.
First, assume that is a point on . If is a zero-dimensional subscheme such that its ideal is contained in , then we get , where the equality is due to Proposition 4.3. We conclude that , which implies . Unless is a point on the line and , we have by Proposition 4.5. Thus we get , as desired.
Now consider the case and is on the line . Suppose that . This means there is a zero-dimensional subscheme of degree such that . Let be the ideal generated by the quadrics in . Comparing Hilbert functions, we see that . By Proposition 4.6(v), we get that and hence is generated by together with one cubic. Thus by Proposition 4.3 the value of the Hilbert function of in degree 3 is . Since has degree , its Hilbert function must be . It follows that the Castelnuovo-Mumford regularity of is two. Hence is generated by quadrics, as is . This implies , which is a contradiction because the Hilbert function of in degree 3 is . This proves . We get equality because by Proposition 4.5.
Second, assume is a point on the line , but not . Since meets only in , the ideal is generated by quadrics (see Proposition 4.6). Hence Proposition 4.7 gives , as above. In the proof of Proposition 4.6(vii), we constructed a Gorenstein scheme of degree supported at one point with the property that . Hence, by definition of the cactus rank, we get . Together with the lower bound this yields . ∎
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 in variables are equal unless with and . In this case we have .
Proposition 4.10.
For any symmetric cubic form , satisfies the SLP.
Proof.
If we are at the cusp of , , , the Hilbert function of is , and clearly has the SLP.
If we are at the flex of i.e. and we have , the Hilbert function of is (Proposition 4.3) and 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 of and , we have (Proposition 4.5), the Hilbert function of is (Proposition 4.3), we easily check that is a basis of and is an SL element since is an isomorphism for .
Assume that is neither the flex nor the cusp of . According to Proposition 4.3 the Hilbert function of is and we distinguish two cases:
Case 1: . In this case, where is the support of the Waring decomposition of . Therefore, any non-zero divisor on is an SL element.
Case 2: . We will show that 3 linear forms are sufficient to provide an SL elements for all of them. To prove that is an SL element for it suffices to show that is injective. The map fails injectivity if there is a linear form such that in , i.e. there is a linear form such that . So, our strategy is to study when the quadratic form drops rank. We first observe that . We consider and we will check when is an SL element for . We claim that quadratic form
drops rank exactly when . Indeed, the matrix
associated to the quadratic form has determinant
Therefore, the quadratic form drops rank exactly when , and we conclude that is an SL element except on the lines and . Finally, we consider the quadratic forms
and
Computing the determinant of the symmetric matrices associated to and we get that the quadratic form and never drop rank simultaneously in points of the lines and . Therefore on these lines either or 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 variables has Waring rank . 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
which gives rise to a rational map from with coordinates to . In order to see that this map is generically onto, we look at the Jacobian
with determinant equal to . This is non-zero except when , which corresponds to the cubic curve and when , which corresponds to the point , i.e., the cusp of .
Proposition 5.1.
The generic symmetric quartic has Waring rank at most and the generic symmetric quintic has Waring rank at most .
Proof.
We start by using the decomposition
which produces a symmetric quartic form of Waring rank at most . We expand this into the basis of symmetric quartics given by to get
Thus we have produced a rational map from with coordinates to . In order to show that this is generically onto, which proves our statement, we consider the Jacobian
which has determinant
This is generically non-zero, showing that the generic symmetric quartic has power sum expansion with at most terms.
For the symmetric quintics, we use the power sum expansion
which gives a rational map from to given by the expansion
The Jacobian of this map is given by
and we get the determinant of that as
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 . ∎
Remark 5.2.
In the case of cubics, Proposition 4.5 shows that the generic rank is indeed . In the case of quartics, we know from the Hilbert function that the generic rank is at least 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 . 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 and our upper bound is higher than this. Here we might be able to use the resolution in order to show that the generic rank is higher than , 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 to described in the proof of Proposition 5.1 we can look at the preimage of . Using the fact that
we get that the preimage is given by the following system of equations
From the first three equations, we get
which gives
For all we can solve this equation with and hence get solutions for , and in terms of . Because of the first equation, we see that when . When , we get that from the first equation and then we can see that the Waring rank of is less than the Waring rank of showing that the solution to our equations give .
Generic rank in the case is which is far larger than 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 -point orbits since such decompositions do not give any contribution to the coefficient of .
For , the generic rank is lower than the bound given by our construction, instead of . Also for , generic rank is lower by one. For , 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.