You are currently viewing a new version of our website. To view the old version click .
Mathematics
  • Feature Paper
  • Article
  • Open Access

27 November 2025

On the Hilbert Function of Partially General Unions of Double Points

Department of Mathematics, University of Trento, 38123 Trento, Italy
The author is a member of Gruppo Nazionale per le Strutture Algebriche e Geometriche e loro Applicazioni of Istituto di Alta Matematica, 00185 Rome, Italy.

Abstract

We study the Hilbert function of a union of a fixed set of double points and a prescribed number of general double points. Our main results are for double points of Veronese varieties and of Segre–Veronese varieties. Among the Segre–Veronese varieties, we study the ones with all factors of dimension one and the ones with two factors, one of them of dimension one. We give many examples with exceptional or controlled behavior for a small number of double points.

1. Introduction

A classical topic of last century algebraic geometry was the study of the secant varieties σ x ( X ) P r of an embedded projective variety X P r [,,]. The key information about them is the dimension of these secant varieties. This topic was revived, because for certain embedded varieties, it gave important results for applied mathematics, roughly speaking related to tensors [,,,,,].
Our main results are about the Veronese embeddings of projective spaces and on the Segre–Veronese varieties. Veronese embeddings of projective spaces are related to the additive decomposition of forms. The dimension of a secant variety of a Segre–Veronese embedding is related to the dimension of the set of all partially symmetric tensors with a fixed format and a fixed partially symmetric tensor rank. Knowing these dimensions is very useful, because a standard way to handle a tensor (or a received packet of data) is to approximate it with a low rank one.
The Terracini Lemma relates this dimension to the dimension of the linear span of a “general” union of double points (ref. [], Cor. 10), i.e., to the following interpolation problem.
For any irreducible quasi-projective variety W, let W reg denote the set of its smooth points. For any positive integer x, let S ( W reg , x ) denote the set of all subsets of W reg with cardinality x. For all p W reg , let ( 2 p , W ) or just 2 p denote the closed subscheme of W with ( I p , W ) 2 as its ideal scheme. The scheme ( 2 p , W ) is a connected zero-dimensional scheme of degree 1 + dim W . For all S S ( W reg , x ) , set ( 2 S , W ) : = p S ( 2 p , W ) . Often we write 2 S if W is clear in a statement. The set S ( W reg , x ) is an irreducible variety of dimension x dim X , and hence, we are allowed to speak about its general element. The Terracini Lemma gives dim σ x ( X ) = dim 2 S for a general S S ( X reg , x ) , where denote the linear span. The case x = 1 of it explains the usefulness of ( 2 p , W ) : if W is embedded in a projective space P r , then ( 2 p , W ) P r is the Zariski tangent space of W at p.
In this paper we start with a fixed A S ( X reg , a ) and give (in some cases) dim 2 A 2 S for a general S S ( X reg , x a ) . The integer dim 2 A 2 S is the dimension of the ( x a ) -secant variety linear projection from the linear space 2 A .
For the Veronese embedding of P n , we prove the following result.
Theorem 1.
Fix positive integers a , x and A S ( P n , a ) . Fix an integer d max { 5 , 2 a 1 } and a general S S ( P n A , x ) . Let V ( n , d ) be the vector space of degree d forms in n + 1 variables. Set V ( n , d ) ( 2 S 2 A ) : = { f V ( n , d ) f | 2 S 2 A = 0 } . Then,
dim V ( n , d ) ( 2 S 2 A ) = max { 0 , n + d n ( n + 1 ) ( a + x ) } .
We will prove Theorem 1 using sheaf cohomology. In terms of sheaf cohomology, Theorem 1 is equivalent to the following statement.
Theorem 2.
Fix positive integers a , x and A S ( P n , a ) . Fix an integer d max { 5 , 2 a 1 } and a general S S ( P n A , x ) . Then, either h 1 ( I 2 S 2 A ( d ) ) = 0 or h 0 ( I 2 S 2 A ( d ) ) = 0 .
It is interesting also to handle 2 S A instead of 2 S 2 A . It corresponds to computing the dimension of a secant variety of the image of X by the linear projection from A (see Proposition 1 for the case of Veronese embeddings).
We use the following fundamental theorem proven by J. Alexander and A. Hirschowitz [,,,].
Theorem 3
(Alexander–Hirschowitz). Fix positive integers n, d and x. Take a general set S S ( P n , x ) . Then, either h 1 ( I 2 S ( d ) ) = 0 or h 0 ( I 2 S ( d ) ) = 0 , except in the following cases:
1. 
n > 1 and d = 2 ;
2. 
( n , d ) { ( 2 , 4 ) , ( 3 , 4 ) , ( 4 , 3 ) , ( 4 , 4 ) } .
This is the sheaf-cohomology version of the Alexander–Hirschowitz theorem. In the language of Theorems, 1 and 3 are stated in the following way.
Theorem 4
(Alexander–Hirschowitz). Fix positive integers n, d and x. Take a general set S S ( P n , x ) . Then, dim V ( n , d ) ( 2 S ) = max { 0 , n + d n x ( n + 1 ) } , except in the following cases:
1. 
n > 1 and d = 2 ;
2. 
( n , d ) { ( 2 , 4 ) , ( 3 , 4 ) , ( 4 , 3 ) , ( 4 , 4 ) } .
Remark 1.
In the statement of Theorem 3, the values of h 0 ( I 2 S ( d ) ) and h 1 ( I 2 S ( d ) ) in the listed exceptional cases were known before [], and they are listed in [,,,], except one case solved in []. There are many proofs of the Alexander–Hirschowitz theorem [,,,].
We recall that the Segre–Veronese varieties are the embeddings by complete linear systems of multiprojective spaces. We study this problem for two types of multiprojective spaces, P m × P 1 and ( P 1 ) k , k 2 .
We use the automorphism group of the multiprojective space to drastically simplify the proofs, reducing to the case of the “worst A”. We think that this reduction is also useful to test computationally for many other interpolation problems. The word “worst” is used, because if this element A of S ( X reg , a ) has good interpolation, then all B S ( X reg , a ) have good interpolation by the semicontinuity theorem for cohomology (ref. [], Th. III.12.8). For subsets of P n , the worst ones are the one collinear (Lemma 3). We are able to describe and use the next worst ones when all but one of the points are collinear (Lemma 4). For the multiprojective spaces and their Segre–Veronese embeddings, there is no single notion of “bad”. We use the one with respect to the first factor (Remark 7). These lemmas and remarks may be used for interpolations with other zero-dimensional schemes, not just double points.
We raise the following interpolation problem.
Question: Fix positive integers a and x. Take an integral and non-degenerate variety X P r and an integral subvariety T X such that T X reg . Compute dim 2 A 2 S and dim ( 2 A , T ) 2 S for a general A S ( T reg X reg , a ) and a general S S ( X reg , x ) .
It may be easier when T is a hypersurface of X. We use it in one case (Lemma 10) to prove all the main theorems of Section 6. The question arises when we use the so-called Differential Horace Lemma according to J. Alexander and A. Hirschowitz (Lemma 2) and an inductive procedure using the hypersurface T. In [], there is a case in which higher codimension is needed and handled.
In Section 2, we fix the notation and give the preliminary remarks and lemmas used in the paper. Most proofs later are inductive, the induction step being handled using in a tricky way the Differential Horace Lemma.
Section 3 contains many lemmas on the Veronese embeddings and a result (Proposition 1) about the Hilbert function of A 2 S , S general in S ( P n , x ) . At the beginning of the section, we describe our method (Remark 6).
Section 4 contains the proof of Theorem 2.
Section 5 and Section 6 are for Segre–Veronese varieties. In Section 5, we give the notation and background and prove the case P m × P 1 (Theorem 5 for 2 A 2 S , Proposition 3 for A 2 S ) using induction on the integer m. Section 6 contains four major results for the embeddings of ( P 1 ) k , k 3 , (Theorems 6–9), the case k = 2 being proven in Section 5. At the beginning of Section 5, we describe the set-up of the Segre–Veronese varieties and give three possible extensions for the interested readers.
Section 7 contains examples for P n and P 1 × P 1 for up to 4 points. We also consider multiple points of multiplicity m > 2 .
In the last section, we propose the following extension of the question: instead of T, we take a flag T 1 T 2 T s X and prescribe the number of points contained in each stratum T i + 1 T i .
We thank the referees for helpful feedback, which led to some changes in the exposition.

2. Preliminary Results

We work over an algebraically closed field K of characteristic 0.
Let M be an integral projective variety, D an effective Cartier divisor of M and Z M a zero-dimensional scheme. The residual scheme Res D ( Z ) of Z with respect to D is the zero-dimensional subscheme of M with I Z , M : I D , M as its ideal sheaf.
We have Res D ( Z ) Z and deg ( Z ) = deg ( Z D ) + deg ( Res D ( Z ) ) . If Z = A B with A B = , then Res D ( Z ) = Res D ( A ) Res D ( B ) . For all line bundles R on M, we have an exact sequence
0 I Res D ( Z ) R ( D ) I Z R I Z D , D R | D 0
For all p M reg , let ( 2 p , M ) denote the zero-dimensional subscheme of M with ( I p , M ) 2 as its ideal sheaf. We have ( 2 p , M ) red = { p } and deg ( ( 2 p , M ) ) = dim M + 1 . For all finite set S M reg , set ( 2 S , M ) : = p S ( 2 p , M ) , with the convention ( 2 , M ) = .
For any quasi-projective variety W and all positive integers x, let S ( W , x ) denote the set of all subsets of W with cardinality x. Set S ( W , 0 ) = { } .
Remark 2.
Let A B P r be zero-dimensional schemes. We have A B and dim B dim A + deg ( B ) deg ( A ) . Thus, if B is linearly independent, then A is linearly independent, while if A = P r , then B = P r .
Remark 3.
Let X P r be an integral and non-degenerate variety. Fix a closed subscheme B X and a positive integer x. Take a general S S ( X B red , x ) . Then we have
dim B S = min { r , dim B + x } .
The following observation is a cohomological version of Remark 3.
Remark 4.
Let X be an integral projective variety, L a line bundle on X, B a closed subscheme of X and x a positive integer. Let S be a general element of S ( X B , x ) . Then
h 0 ( I B S L ) = max { 0 , h 0 ( I B L ) x } ,
h 1 ( I B S L ) = h 1 ( I B L ) + max { 0 , x h 0 ( I B L ) } .
Remark 5.
Fix an integral and non-degenerate variety X P r and a linear subspace M P r . Set m : = dim M . If m = r 1 , then M { p } = P r . Now assume m r 2 . Let l M : P r M P r m 1 denote the linear projection from M. Since X is non-degenerate, X M . Hence, μ : = l M | X M X M P r m + 1 is a morphism. Since X is non-degenerate, μ ( X M X ) spans P r m 1 . Hence, μ is not a constant map. Since we are in characteristic zero and dim X 1 , the differential of μ is non-zero at a general p X reg . Hence dim M v = m + 2 for a general connected zero-dimensional subscheme v of X reg such that deg ( v ) = 2 .
Lemma 1.
Fix an integral projective variety X of positive dimension, L a line bundle on X and D an effective irreducible Cartier divisor of X such that h 1 ( L ( D ) ) = 0 . Let Z X be a zero-dimensional scheme such that h 1 ( I Z L ) = 0 . Set a : = h 0 ( I Z L ) . Take a positive integer x and a general S S ( D , x ) . Assume h 0 ( I Res D ( Z ) L ( D ) ) max { 0 , a x } . Then h 0 ( I Z S L ) = max { 0 , a x } and h 1 ( I Z S L ) = max { 0 , x a } .
Proof. 
Remark 4 gives h 1 ( L ) = 0 . Take a maximal S S with the property that we have h 0 ( I Z S L ) = max { 0 , a # S } (we allow the case S = ). Assume S S (and hence, assume h 0 ( I Z S L ) > 0 ) and take p S S . Note that Res D ( Z S ) = Res D ( Z ) . Since p is general in D, the maximality of S implies that D is the base locus of | I Z S L | . Hence, we obtain h 0 ( I Res D ( Z ) L ( D ) ) = a # S , a contradiction.   □
The following result according to J. Alexander and A. Hirschowitz is called the Differential Horace Lemma. It was proven and used in this generality in [] and in their previous joint papers.
Lemma 2
(Differential Horace Lemma). Let M be an integral projective variety, R a line bundle on M, Z M a zero-dimensional scheme and D an integral effective Cartier divisor. Fix an integer x > 0 , a general S S ( M , x ) and a general A S ( D , x ) . Then for all i N , we have
h i ( M , I Z ( 2 S , M ) R ) h i ( I Res D ( Z ) A R | D ) + h i ( I Z ( 2 S , M ) R ( D ) )
The case S = of Lemma 2 is called the Horace Lemma.

3. Veronese Embeddings

Remark 6.
The Alexander–Hirschowitz theorem was proven by double induction on the dimension, n, of the projective space and the degree, d, of the forms considered for the interpolation problem “x general double points”. The main difficulty is that the cases with very low d are exceptional. If for some d, one knows the theorem in P n 1 and P n for the degree d, then it would be not too difficult to prove it in P n for forms of degree d + 1 . For our interpolation problems in our main theorems we are not (by our choice) in the exceptional range. We show some exceptional cases in Section 7. Section 7 shows that our tools work also to discover exceptional cases and prove their exact defect (the integer h 1 ( I 2 A 2 S ( d ) ) ). In Section 3 and Section 4, we reduce the case of 2 A 2 S for d to a case for 2 S , S general for a lower d such that ( n , d , # S ) is not in the list of exceptional cases. It is a key part of a less than 80 page long manuscript to use as much as possible the statement of the Alexander–Hirschowitz theorem. We first show that instead of an arbitrary A S ( P n , a ) , it is sufficient to handle all B S ( P n , a ) contained in a line L: if 2 B 2 S is not exceptional in degree d, then 2 A 2 S is not exceptional in degree d (Lemmas 3 and 4). Points B contained in a line L are easy to control, because for the Differential Horace Lemma, we take a hyperplane H L , and in the residual, the multiplicity of each point decreases by 1. Hence, in 2 steps, we go from 2 B to ∅. This observation may be used for many other interpolation problems. Instead of 2 A , we may take multiple points of arbitrary multiplicity in the sense of Notation 1. See [] for the Differential Horace Lemma for multiple points of multiplicity m > 2 . Our proofs are in some sense from d to a lower d . Instead of the set (A or B) of cardinality a which we had at level d, we have the set ∅ at level d . To do that (with # S as small as needed), we use not only the Differential Horace Lemma but Lemma 1, which drastically reduces the degree of the zero-dimensional scheme that we need to control in degree d 1 and d 2 . See [,,] for some uses of this approach to the Differential Horace Lemma, refs. [,] not being restricted to the case of Segre–Veronese varieties. The case A 2 S is carried out in Proposition 1.
See (ref. [] Figure 11, p. 260) as a much more complicated example of linear projection as a flat limit of a family of automorphisms. This example motivated the following two lemmas.
Lemma 3.
Fix positive integers a, x, d and A S ( P n , a ) . Then, there is B S ( P n , x ) such that B is contained in a line, and for a general S S ( X A , x ) , we have:
1. 
h 1 ( I 2 A 2 S ( d ) ) h 1 ( I 2 B 2 S ( d ) ) ;
2. 
h 0 ( I 2 A 2 S ( d ) ) h 0 ( I 2 B 2 S ( d ) ) ;
3. 
h 1 ( I A 2 S ( d ) ) h 1 ( I 2 B 2 S ( d ) ) ;
4. 
h 0 ( I A 2 S ( d ) ) h 0 ( I 2 B 2 S ( d ) ) .
Proof. 
Projectively equivalent zero-dimensional schemes have the same Hilbert function, i.e., for all d N , the vector spaces of degree d forms vanishing on them have the same dimension. By the semicontinuity theorem for cohomology (ref. [], Th. III.12.8), it is sufficient to prove that there is B S ( L , a ) , which is a flat limit of a family of elements of S ( P n , a ) , all of them projectively equivalent to B. Fix a line L P n . Take a general codimension 2 linear subspace V of P n . Since V is general, A V = , L V = and L V = P n . Thus, we may take a homogeneous system of coordinates x 0 , , x n such that V = { x 0 = x 1 = 0 } and L = { x 2 = = x n = 0 } . Let h : P n V P 1 denote the linear projection for V. We identify L with the target of h. Since V is general, V does not intersect any line spanned by 2 points of A. Hence, # h ( A ) = a . Set B : = h ( A ) S ( L , a ) . For any non-zero constant c, let h c denote the linear automorphism of P n defined by the formula h c ( [ x 0 : x 1 : x 2 : : x n ] ) = [ x 0 : x 1 : c x 2 : : c x n ] . Each h c ( A ) is projectively equivalent to A. If we take c = 0 in the formula for h c , we obtain h. Since # h ( A ) = a , B is the flat limit of the family { h c ( A ) } c 0 .   □
Lemma 4.
Take a finite set S P n , n 2 , p S and a line L P n such that p L . Set A : = S { p } . Assume that no line spanned by 2 points of A contains p. Then, there is B L such that # B = # S , and { p } B is the flat limit of a family of subsets of P n projectively equivalent to S.
Proof. 
Take a general codimension 2 linear subspace V of P n containing p.
Let h : P n V P 1 denote the linear projection for V. We identify L with the target of h. Take a homogeneous system of coordinates x 0 , , x n such that V = { x 0 = x 1 = 0 } , p = [ 0 : 0 : : 1 ] and L = { x 2 = = x n = 0 } . Set B : = h ( A ) . Since no line spanned by 2 points of A contains p, and V is a general codimension 2 linear space containing p, V does not intersect any line spanned by 2 points of A. Hence, # B = # A . For any non-zero constant c, let h c denote the linear automorphism of P n defined by the formula h c ( [ x 0 : x 1 : x 2 : : x n ] ) = [ x 0 : x 1 : c x 2 : : c x n ] . Each h c ( A ) is projectively equivalent to B. Note that if we take c = 0 in the definition of the automorphism h c , we obtain the linear projection h. Since # h ( A ) = a , B is the flat limit of the family { h c ( A ) } c 0 . Since h c ( p ) = p , { p } B is the flat limit of the family { { p } h c ( A ) } c 0 , concluding the proof.   □
Proposition 1.
Fix integers a 3 , n 2 and d max { a 1 , 5 } . Take A S ( P n , a ) and a general S S ( P 2 A , x ) . If ( n , d , a ) = ( 2 , 4 , 5 ) , assume x 5 . Then, either h 1 ( I 2 S A ( d ) ) = 0 or h 0 ( I 2 S A ( d ) ) = 0 .
Proof. 
By Lemma 3, it is sufficient to prove the case in which A is contained in a line L. We will prove (inductively) a stronger result in which we allow as A any degree a zero-dimensional subscheme of L.   □
Set y t , n : = ( n + t n a ) / ( n + 1 ) and x t , n : = ( n + d n a ) / ( n + 1 ) , where t is a positive integer. By Remark 2 to prove Proposition 1 for A and d and all x, it is sufficient to prove that h 1 ( I 2 A 2 E ( d ) ) = 0 for a general E S ( P n A , y d , n ) , and h 0 ( I A 2 F ( d ) ) = 0 for a general F S ( P n A , x d , n ) .
Take z d , n { y d , x d , n } and a general S S ( P 2 L , z d , n ) .
(a) Assume n = 2 . First, assume d = a 1 . Recall that S is general. Since h i ( L , I A , L ( a 1 ) ) = 0 , i = 0 , 1 , Res L ( A ) = and S L = , the residual exact sequence of L gives h i ( I 2 S A ( d ) ) = h i ( I 2 S ( d 1 ) ) , i = 0 , 1 . Since we excluded the case ( d , x ) = ( 4 , 5 ) , it is sufficient to use the Alexander–Hirschowitz theorem (Theorem 3). Now assume d a and that for all a a , the proposition is true for the integer d 1 .
(a1) Assume d a odd. Since 3 y d , 2 d + 2 2 a 2 and d a + 1 , we have the inequality y d , 2 ( d + 1 a ) / 2 . We specialize S to S 1 S 2 with S 2 a general element of S ( L , ( d + 1 a ) / 2 ) and S 2 a general element of S ( P 2 , z d , 2 ( d + 1 ) / 2 ) . Note that we have L ( A 2 S 1 2 S 2 ) = A ( 2 S 2 , L ) and Res L ( A 2 S 1 2 S 2 ) = 2 S 1 S 2 . Since h i ( L , I A ( 2 S 2 , L ) ( d ) ) = 0 , i = 0 , 1 , the residual exact sequence of L shows that to prove the proposition in degree d, it is sufficient to prove that either h 1 ( I 2 S 1 S 2 ( d 1 ) ) = 0 (case z d , 2 = y d , 2 ) or h 0 ( I 2 S 1 S 2 ( d 1 ) ) = 0 (case z d , 2 = x d , 2 ). Since # S 2 = ( d 1 ) / 2 d 1 , we may use the inductive assumption, unless d 1 = 4 . Assume d = 5 . Since d a is odd, we obtain a = 4 , # S 2 = 1 , y 5 , 2 = 5 and x 5 , 2 = 6 . Take the case z 5 , 2 = y 5 , 2 . We have h 1 ( I 2 S 1 ( 4 ) ) = 0 by the Alexander–Hirschowitz theorem, and hence, h 0 ( I 2 S 1 ( 4 ) ) = 3 . Since P 2 is homogeneous and we may take S 2 as general after fixing S 1 , we obtain h 0 ( I 2 S 1 S 2 ( 4 ) ) = 2 , i.e., h 1 ( I 2 S 1 S 2 ( 4 ) ) = 0 . Now, assume z 5 , 2 = x 5 , 2 . By the Alexander–Hirschowitz theorem, we have h 1 ( I 2 S 1 ( 4 ) ) = h 0 ( I 2 S 1 ( 4 ) ) = 1 with | I 2 S 1 ( 4 ) | being the double of the conic containing S 1 . Hence, h 0 ( I 2 S 2 S 1 ( 4 ) ) = 0 for a general S 2 .
(a2) Assume d a even. Fix a general B 1 L such that # B 2 = ( d a ) / 2 , a general p L ( A B 1 ) and a general B 2 S ( P 2 , z d , 2 1 ( d a ) / 2 ) . Note that h i ( L , I A ( 2 B 2 , L ) { p } , L ( d ) ) = 0 , i = 0 , 1 .
By the Differential Horace Lemma (Lemma 2), to prove the theorem in degree d, it is sufficient to prove that either h 1 ( I 2 B 1 B 2 ( 2 p , L ) ( d 1 ) ) = 0 (case z d , 2 = y d , 2 ) or h 0 ( I 2 B 1 B 2 ( 2 p , L ) ( d 1 ) ) = 0 (case z d , 2 = x d , 2 ). It is for using induction on d that we allowed an arbitrary zero-dimensional scheme contained in a line. We have deg ( B 2 + ( 2 p , L ) ) = ( d a ) / 2 + 2 d . Hence, we may use the inductive assumption, unless d = 5 . Since a 3 and d a is even, we obtain a = 3 , and hence, y 5 , 2 = x 5 , 2 = 6 . We use a different proof but first note that the proof just given for the general case would give h 1 ( I A 2 S ( 5 ) ) = h 0 ( I A 2 S ( 5 ) ) 1 . Take a smooth conic C such that deg ( C A ) = 1 . We specialize S to the union E 1 { q } with E 1 , the union of 5 general points of C and q general in P 2 . Since C P 1 and h 0 ( C , O C ( 5 ) ) = 11 , we obtain h i ( C , I ( 2 E 1 , C ) A C , C ( 5 ) ) = 0 , i = 0 , 1 . Note that h 0 ( I E 1 2 q Res L ( A ) ( 3 ) ) = h 1 ( I E 1 2 q Res L ( A ) ( 3 ) ) . Thus, by the residual exact sequence of C, it is sufficient to prove that h 0 ( I E 1 2 q Res L ( A ) ( 3 ) ) = 0 . Assume h 0 ( I E 1 2 q Res L ( A ) ( 3 ) ) = 1 . Since deg ( C A ) = 1 , deg ( C Res L ( A ) ) 1 . The residual exact sequence of C gives h 1 ( I E 1 Res L ( A ) ( 3 ) ) = 0 . Thus, h 0 ( I E 1 Res L ( A ) ( 3 ) ) = 3 . The linear system | I E 1 Res L ( A ) ( 3 ) | has dimension 2, and all its elements are singular at q. Since q may be a general point of P 2 , the Bertini theorem gives a contradiction.
(b) Assume n > 2 and that the proposition is true in lower dimensional projective spaces. Take a hyperplane H L . In step (b1), we take z d , n = y d , n . In step (b2), we take z d , n = y d , n . Set e d , n 1 : = n + d 1 n 1 n y d , n 1 . Note that 0 e d , n 1 n 2 . We use the following numerical claims.
Claim 1.
We have y d , n y d , n 1 + n 1 .
Proof of Claim 1.
Assume y d , n y d , n 1 + n 2 . Multiplying this inequality by ( n + 1 ) n and using that ( n + 1 ) y d , n n + d n n and n y d , n 1 n + d n , we obtain
n n + d n n 2 ( n + 1 ) n + d 1 n 1 + n ( n + 1 ) ( n 2 )
We have n n + d n ( n + 1 ) n + d 1 n 1 = ( d 1 ) n + d 1 n 1 . Hence, (2) is equivalent to
( d 1 ) n + d 1 n 1 n 3 2 n
The left hand side of (3) is an increasing function of d. Thus, to prove Claim 1 by contradiction, it is sufficient to prove that (3) fails for d = 5 . Take d = 5 . The left hand side of (3) is 4 ( n + 4 ) ( n + 3 ) ( n + 2 ) ( n + 1 ) > n 3 2 n for all n 3 .   □
Claim 2.
We have y d , n y d , n 1 y d 1 , n 1 .
Proof of Claim 2.
Assume y d , n y d , n 1 + y d 1 , n . We multiply this inequality by n ( n + 1 ) and then use that n ( n + 1 ) y d , n n n + d n ,
n ( n + 1 ) y d , n 1 ( n + 1 ) n + d 1 n 1 ( n 1 ) ( n + 1 ) ,
and n ( n + 1 ) y d 1 , n n n + d 1 n n 2 . We have n + d n n + d 1 n 1 = n + d 1 n 1 . We obtain the inequality
n + d 1 n 1 ( n 1 ) ( n + 1 ) + 2 n 2 n 1
The left hand side of this inequality is an increasing function of d. Since d 5 and n 3 (4) gives ( n + 4 ) ( n + 3 ) ( n + 1 ) n 3 n 2 n 1 . This contradiction proves Claim 2.
(b1) Take z n , d = y n , d . Claim 1 gives y d , n y d , n 1 + e d , n 1 . Take a general set S 1 S ( P n , y d , n y d , n 1 e d , n 1 ) and a general ( S 2 , S 3 ) S ( H , y d , n 1 ) × S ( H , e d , n 1 ) . Since Proposition 1 is proven for H, Remarks 3 and 4 give h i ( H , I ( 2 S 2 , H ) S 3 , H ( d ) ) = 0 , i = 0 , 1 . Hence, the Differential Horace Lemma (Lemma 2) shows that to prove the proposition for n, d and x : = y d , n 1 , it is sufficient to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , H ) ( d 1 ) ) = 0 . Note that deg ( 2 S 1 S 3 ( 2 S 2 , H ) ) = n + d 1 n ( n + d n ( n + 1 ) y d , n ) n + d 1 n .   □
Claim 3.
We have h 1 ( I 2 S 1 ( 2 S 3 , H ) ( d 1 ) ) = 0 .
Proof of Claim 3.
Since ( 2 S 3 , H ) 2 S 3 , Remark 2 gives that it is sufficient to prove that h 1 ( I 2 S 1 2 S 3 ( d 1 ) ) = 0 . Note that S 1 S 3 does not depend on A. Since e d , n 1 n 1 and any n 1 points of P n are contained in a hyperplane, h 1 ( I 2 S 1 2 S 2 ( d 1 ) ) = 0 if and only if h 1 ( I 2 E ( d 1 ) ) = 0 for a general E S ( P n , y d , n y d , n 1 ) . Since y d , n y d , n 1 y d 1 , n (Claim 2) and d 5 , it is sufficient to quote the Alexander–Hirschowitz theorem, unless d = 5 and 3 n 4 . In these cases, we use that y d , n y d , n 1 y d 1 , n 1 (Claim 2).   □
We have Res H ( 2 S 1 S 2 ( 2 S 3 , H ) ) = 2 S 1 and deg ( 2 S 1 S 3 ( 2 S 2 , H ) y d 1 , n . Using Claim 3 and Lemma 1 to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , H ) ( d 1 ) ) = 0 , it is sufficient to prove that h 0 ( I 2 S 1 ( d 2 ) ) max { 0 , n + d 1 n deg ( 2 S 1 ) y d , n 1 n e d , n 1 } . Assume for the moment that ( n , d ) ( 4 , 5 ) . Since ( n , d 2 ) is not in the Alexander–Hirschowitz list of exceptional cases, h 0 ( I 2 S 1 ( d 2 ) ) = max { 0 , n + d 2 n deg ( 2 S 1 ) } . Since
n + d 1 n n + d 2 n = n + d 2 n 1 ,
for ( n , d ) ( 4 , 5 ) , it is sufficient to check the inequality y d , n 1 + e d , n 1 n + d 2 n 1 . We have n y d , n 1 + e d , n 1 = n + d 1 n 1 and
n n + d 2 n 1 n + d 1 n 1 = ( n + d 2 ) ! ( n 1 ) ! d ! ( n d ( d 2 ) n d 1 ) .
Thus, it is sufficient to check the inequality ( n 1 ) e d , n 1 ( n + d 2 ) ! ( n 1 ) ! d ! ( n d ( d 2 ) n d 1 ) . Since e d , n 1 n 1 , it is sufficient to check the inequality
( n 1 ) 2 ( n + d 2 ) ! ( n 1 ) ! d ! ( n d ( d 2 ) n d 1 )
Since n + d 1 d + 1 , the right hand side of (5) is an increasing function of d. For d = 5 , the inequality (5) is the inequality
( n 1 ) 2 ( n + 3 ) ( n + 2 ) ( n + 1 ) n ( 14 n 6 ) / 120 ,
which is true for all n 3 .
Now, assume n = 4 and d = 5 . Since d a 1 , we obtain 3 a 6 . We have y 5 , 4 = 24 , x 5 , 4 = 25 . If 3 a 4 , then y 5 , 3 = 13 and e 5 , 3 = 4 a . If 5 a 6 , then y 5 , 3 = 12 and e 5 , 3 = 8 a . It is sufficient that for a general E S ( P 4 , 7 ) , we have h 0 ( I 2 E ( 3 ) ) = 1 by the Alexander–Hirschowitz theorem.
(b2) Now, assume z d , n = x d , n . By step (b1), we may assume y d , n < z d , n . Thus, we have z d , n = y d , n + 1 . We specialize S to S 3 S 4 with S 3 a general element of S ( P n , y d , n y d , n 1 ) and S 4 a general union of y d , n 1 + 1 points of H. We have y d , n 1 + 1 x d , n 1 . The inductive assumption gives h 0 ( H , I A ( 2 S 4 , H ) ( d ) ) = 0 . Hence, to prove Proposition 1 in this case, it is sufficient to prove that h 0 ( I 2 S 3 S 4 ( d 1 ) ) = 0 . Since d 1 > 2 , the the triple ( n , d 1 , # S 3 ) is not in the Alexander–Hirschowitz list, and hence,
h 0 ( I 2 S 3 ( d 1 ) ) = max { 0 , d + n n ( n + 1 ) ( y d , n y d , n 1 ) } .
Then, we use Lemma 1.

4. Proof of Theorem 2

In this section, we prove Theorem 2.
For all integers a, n and d such that a 0 , n 1 and d a 1 , set
y d , n , a : = ( n + d n a ) / ( n + 1 ) , x d , n , a : = ( n + d n a ) / ( n + 1 )
and e d , n , a : = n + d n a ( n + 1 ) y d , n , a . We have y d , n , a x d , n , a y d , n , a + 1 , 0 e d , n , a n , ( n + 1 ) y d , n , a + a + e d , n , a . Thus, we have y d , 1 , a = ( d + 1 a ) / 2 , e d , 1 , a = 0 if d a is odd and e d , 1 , a = 1 if d a is even.
For all integers a, n and d such that a 0 , n 1 and d 2 a 1 , set
u d , n , a : = n + d n / ( n + 1 ) a , v d , n , a : = n + d n / ( n + 1 ) a and
f d , n , a : = n + d n ( n + 1 ) a ( n + 1 ) u d , n , a .
We have u d , n , a v d , n , a u d , n , a + 1 , 0 f d , n , a n , ( n + 1 ) u d , n , a + ( n + 1 ) a + f d , n , a . Thus, we have y d , 1 , a = ( d + 1 a ) / 2 , f d , 1 , a = 0 if d is odd, f d , 1 , a = 1 if d is even, while we have u d , n , a = y n , d , 0 a = u n , d , 0 a , v d , n , a = v n , d , 0 a = v n , d , 0 a and f d , n , a = f d , n , 0 .
If a is fixed (in a statement or a proof), then we write y n , d , x n , d , v n , d , e n , d and f n , d instead of y n , d , a , x n , d , a , v n , d , a , e n , d , a and f n , d , a . For a fixed a, the integers y n , d and x n , d are the ones used in the proof of Lemma 1. We fix a and ϵ { 1 , 2 } . For all n 1 , we use the integer y n , d ( ϵ ) : = y n , d ϵ , x n , d ( ϵ ) : = x n , d ϵ and e d , n ( ϵ ) : = e d , n .
Remark 7.
Let X P r be an integral and non-degenerate surface. Fix a linear space M P r . Then dim M v = min { r , 2 + dim M } for a general connected degree 2 zero-dimensional scheme whose reduction is contained in X reg (Remark 5).
Remark 8.
Fix ( n , d , a ) and assume f d , n , a 2 . By Remark 7, to prove Theorem 2 for the triple ( n , d , a ) , it is sufficient to prove it for x = u d , n .
Remark 9.
Assume a = 3 . Take a line L P n , A L such that # A = 3 and a hyperplane H L . By Remark 8, if f d , n 2 to prove the case for this triple ( n , d , A ) , it is sufficient to check the case x = u d , n . Take a general S S ( P n , u d , n ) .
(a) Take n = 2 and d = 5 . We have u 5 , 2 = 4 and f 5 , 2 = 0 . Since h i ( L , I ( 2 A , L ) , L ( 5 ) ) = 0 , i = 0 , 1 , to prove Theorem 2 in this case, it is sufficient to use that h 1 ( I 2 S A ( 4 ) ) = 0 (Lemma 1).
(b) Take n = 2 and d = 6 . We have u 6 , 2 = 6 and f 6 , 2 = 1 . Take a smooth plane cubic C such that C L = A and degenerate S to a general S S ( C , 6 ) . Note that we have deg ( ( 2 S 2 A ) C ) ) = 18 = deg ( O C ( 6 ) ) . For a general S , the scheme ( 2 S 2 A ) C is not the complete intersection of C with a degree 6 plane curve. Since C is an elliptic curve, h i ( C , I ( 2 S 2 A ) C , C ( 6 ) ) = 0 , i = 0 , 1 . By the residual exact sequence of C, it is sufficient to prove that h 0 ( I A S 1 ( 3 ) ) = 0 . This is is true according to the Bezout theorem and the fact that for a general S 1 the set A S 1 is not a complete intersection of C with another plane cubic.
(c) Take n = 3 and d = 5 . We have u 5 , 3 = 11 and f 5 , 3 = 0 . We degenerate S to S 1 S 2 with S 1 general in S ( P 3 , 5 ) and S 2 general in S ( H , 4 ) . Part (a) gives h i ( H , I ( 2 A , H ) ( 2 S 2 , H ( 5 ) ) = 0 . Hence, it is sufficient to prove that h i ( I 2 S 1 A S 2 ( 4 ) ) = 0 . Proposition 1 gives h 1 ( I 2 S 1 A ( 4 ) ) = 0 . By Lemma 1, it is sufficient to use that h i ( I 2 S 1 ( 3 ) ) = 0 , i = 0 , 1 .
(d) Take n = 3 and d = 6 . We have u 6 , 3 = 18 and f 6 , 3 = 0 . We take a general S S ( P 3 , 11 ) a general S S ( H , 6 ) and a general q H . Step (b) gives the equalities h i ( H , I ( 2 S , H ) ( 2 A , H ) { q } , H ( 6 ) ) = 0 , i = 0 , 1 . Hence, the Differential Horace Lemma (Lemma 2) applied to q shows that it is sufficient to prove the quality h i ( I 2 S A S ( 2 q , H ) ( 5 ) ) = 0 . Since q is general in a general plane containing L, S { q } may be seen as a general element of S ( P 3 , 12 ) . Proposition 1 gives h 1 ( I 2 S 2 q A ( 5 ) ) = 0 . Hence, h 1 ( I 2 S ( 2 q , H ) A ( 5 ) ) = 0 (Remark 2). Since S is general in H, to conclude this case, it is sufficient to observe that h 0 ( I 2 S ( 4 ) ) = 0 (use for x 9 the case ( n , d , x ) = ( 3 , 4 , x ) is not defective).
(e) Take n = 4 and d = 5 . We have u 5 , 4 = 22 and f 5 , 4 = 1 . We degenerate S to S S with S general in S ( P 4 , 11 ) and S general in S ( H , 11 ) . Part (c) gives the equalities h i ( H , I ( 2 A , H ) ( 2 S , H ) , H ( 5 ) ) = 0 , i = 0 , 1 . Hence, it is sufficient to prove the equality h 1 ( I 2 S A S ( 4 ) ) = 0 . Proposition 3 gives h 1 ( I 2 S A ( 4 ) ) = 0 , and hence, h 0 ( I 2 S A ( 4 ) ) = 12 . By Lemma 1, it is sufficient to use that h 0 ( I 2 S ( 3 ) ) = 0 .
(f) Take n = 4 and d = 6 . We have u 6 , 4 = 39 and f 6 , 4 = 0 . We take a general S S ( P 4 , 21 ) and a general S S ( H , 18 ) . Part (d) gives h i ( H , I ( 2 A , H ) ( 2 S , H ) , H ( 6 ) ) = 0 . By the Horace Lemma, it is sufficient to prove that h i ( I 2 S A S ( 5 ) ) = 0 . Proposition 1 gives h 1 ( I 2 S A ( 5 ) ) = 0 , i.e., h 0 ( I 2 S A ( 5 ) ) = 18 . Using Lemma 1 to conclude, it is sufficient to use that h 0 ( I 2 S ( 4 ) ) = 0 .
Lemma 5.
Assume a 3 and d 2 a 1 . Then,
( n + 1 ) ( v d , n u d , n 1 f d , n 1 + f 1 ) n + d 2 n .
Proof. 
We have ( n + 1 ) v d , n 1 n + d n a + n . Since n u d , n 1 + f d , n 1 = n + d 1 n 1 a and 0 f d , n 1 n 1 , n ( u d , n 1 + f d , n 1 ) n + d 1 n 1 . Note that n n + d n n n + d 1 n 1 = n n + d 1 n and that n n + d 1 n = d n + d 1 n 1 ( d 1 ) . Hence, to prove that
n ( n + 1 ) ( v d , n u d , n 1 f d , n 1 + f 1 ) n n + d 2 n ,
it is sufficient to observe that ( d 1 ) n + d 1 n 1 n 2 + a for all d max { 5 , 2 a 1 } and all n 2 .   □
Proof of Theorem 2.
Since any 2 points of P n are projectively equivalent, the Alexander–Hirschowitz theorem (Theorem 3) shows that it is sufficient to use the case a 3 , assuming d 2 a 1 5 . The integer a is fixed, and hence, in this proof, u d , n : = u d , n , a , and so on.
By Lemma 3, we may assume that A is contained in a line L. To prove the theorem for fixed A, n and d and all positive integers x, it is sufficient to prove that h 1 ( I 2 A 2 E ( d ) ) = 0 for a general E S ( P n A , u d , n ) and h 0 ( I 2 A 2 F ( d ) ) = 0 for a general F S ( P n A , v d , n ) . Fix a hyperplane H L . Take z d , n { u d , n , v d , n } and a general S S ( P n , z d n ) . Since S is general, S H = . Note that the theorem is true for n = 1 . Hence, even in the case n = 2 , we may assume that the theorem is true in lower dimensional projective spaces. Hence, we may apply Proposition 3 and the theorem in H, even if n = 2 , and hence, H = L .
Fix an integer d 2 a 1 . Recall that 0 f d , n 1 n 1 . Until step (d), we assume z d , n u d , n 1 + f d , n 1 . Set f : = f d , n 1 .
Take a general ( S 1 , S 2 , S 3 ) S ( P n , z d , n u d , n 1 f ) × S ( H , u d , n 1 ) × S ( H , f ) .
By the inductive assumption on the dimension of the projective space, we have h j ( H , I ( 2 A , H ) ( 2 S 2 , H ) S 3 , H ( d ) ) = 0 , j = 0 , 1 . By the Differential Horace Lemma, to conclude the proof in this case, it is sufficient to prove that h i ( I 2 S 1 S 2 A ( 2 S 3 , H ) ( d 1 ) ) = 0 .
(a) In this step, we prove that h 1 ( I 2 S 1 A ( 2 S 3 , H ) ( d 1 ) ) = 0 .
Set e 1 : = ( ( d + n 2 n 1 ) a u d , n 1 n f ) / n and f 1 : = n + d 2 n 1 a u d , n 1 n e 1 . We have 0 f 1 n 1 .
We assume for the moment that z d , n u d , n 1 + f + e 1 + f 1 , leaving the case z d , n < u d , n 1 + f + e 1 + f 1 to step (c). Take a general ( S 4 , S 5 ) S ( H , e 1 ) × S ( H , f 1 ) . We specialize e 1 of the points of S 1 to different points of S 4 and apply the Differential Horace Lemma (Lemma 2) to the points of S 5 . By the inductive assumption, we have h j ( H , I A ( 2 S 3 , H ) ( 2 S 4 , H ) S 5 , H ( d 1 ) ) = 0 , j = 0 , 1 . By the Differential Horace Lemma, to prove that h 1 ( I 2 S 1 A ( 2 S 3 , H ) ( d 1 ) ) = 0 , it is sufficient to prove that h 1 ( I 2 S 1 S 4 ( 2 S 5 , H ) ( d 2 ) ) = 0 .
(a1) In this step, we prove that h 1 ( I 2 S 1 ( 2 S 5 , H ) ( d 1 ) ) = 0 . By Remark 2, it is sufficient to prove that h 1 ( I 2 S 1 2 S 5 ( d 2 ) ) = 0 . We have # ( S 1 S 5 ) = z d , n u d , n 1 f + f 1 . Assume for the moment that ( n , d 2 , # ( S 1 S 5 ) ) is not in the list of the exceptional cases of the Alexander–Hirschowitz theorem (Theorem 3). Since z d , n v n , d , to prove that h 1 ( I 2 S 1 ( 2 S 5 , H ) ( d 1 ) ) = 0 , it is sufficient to use that Lemma 5 gives the inequality
( n + 1 ) ( v d , n u d , n 1 f + f 1 ) n + d 1 n .
Now, assume that ( n , d 2 , # ( S 1 S 5 ) ) is in the list of the exceptional cases of the Alexander–Hirschowitz theorem. Since d 2 a 1 with a 3 , we obtain a = 3 and
( n , d ) { ( 2 , 6 ) , ( 3 , 6 ) , ( 4 , 5 ) , ( 4 , 6 ) } .
These cases are carried out in Remark 9.
(a2) In this step, we prove that h i ( I 2 S 1 A ( 2 S 3 , H ) ( d 1 ) ) = 0 . Step (a1) gives h 0 ( I 2 S 1 A ( 2 S 3 , H ) ( d 1 ) ) = n + d 1 n deg ( 2 S 1 ) a n f . By Lemma 1 and step (a1), to prove the theorem in this case, it is sufficient to prove that
h 0 ( I 2 S 1 ( d 2 ) ) max { 0 , n + d 1 n deg ( 2 S 1 ) a n f u d , n 1 } .
Assume for the moment that ( n , d 2 , # S 1 ) is not in one of the exceptional cases of the Alexander–Hirschowitz theorem. We have h 0 ( I 2 S 1 ( d 2 ) ) = n + d 2 n deg ( 2 S 1 ) . Hence, it is sufficient to use that f n and n + d 2 n 1 a + n 2 .
Since d 5 , the exceptional cases for ( n , d 2 ) are settled in Remark 9.
(b) In this step, we conclude the proof of Theorem 2, proving the equality h i ( I 2 S 1 S 2 ( 2 S 3 , H ) A ( d 1 ) ) = 0 . Step (a) gives
h 1 ( I 2 S 1 ( 2 S 3 , H ) ( d 1 ) ) = n + d 1 n deg ( 2 S 1 ) a n f .
Since S 2 is general in H and # S 2 = u d , n 1 , Lemma 1 gives that it is sufficient to prove that h 0 ( I 2 S 1 ( d 2 ) ) max { 0 , n + d 1 n deg ( 2 S 1 ) a n f u d , n 1 } . Remember that d 2 a 1 5 . Hence, if ( n , d 2 ) is in the exceptional list of the Alexander–Hirschowitz theorem, we use Remark 9. If ( n , d 2 ) is not in this list, then h 0 ( I 2 S 1 ( d 2 ) ) = max { 0 , n + d 2 n deg ( 2 S 1 ) } . Hence, it is sufficient to prove that a + n f + u d , n 1 n + d 2 n 1 . We have n ( u n , d 1 + f ) = n + d 1 n 1 n a . Hence, it is sufficient to prove that a + ( n 2 1 ) f n n + d 2 n 1 n + d 1 n 1 . Since f ( n 1 ) , it is sufficient to prove that a + ( n 2 1 ) ( n 1 ) n n + d 2 n 1 n + d 1 n 1 . Call t ( n , d , a ) the difference between the right hand side and the left one of the inequality we want to prove. We have t ( 2 , d , a ) = d 1 a 3 , which is 0 , except in the case a = 3 , n = 2 , d 6 settled in Remark 9. We have t ( 3 , d , a ) = d 2 1 a 16 . Since a 3 and d 2 a 1 , t ( 3 , d , a ) > 0 , we have t ( 4 , d , a ) = ( d + 2 ) ( d + 1 ) ( d 1 ) / 2 16 a > 0 . Call w ( n , d , a ) the left hand side of the inequality we want to prove, and q ( n , d , a ) its right hand side. We have q ( n , d , a ) = ( n + d 2 ) ! ( n d n d + 1 ) ( n 1 ) ! d ! . We have w ( n + 1 , d , a ) / w ( n , d , a ) ( n + 1 ) 3 / n 3 . We have q ( n + 1 , d , a ) / q ( n , d , a ) = ( n + d 1 ) ( ( n + 1 ) d n d ) n ( n d n d + 1 ) . We have ( ( n + 1 ) d n d ) = ( n 1 ) d . For a fixed n, the rational number q ( n + 1 , d , a ) / q ( n , d , a ) increases with d. Hence, it is sufficient to use that q ( n + 1 , 5 , a ) / q ( n , 5 , a ) > 5 / 4 .
(c) Assume e + f z d , n < e + f + e 1 + f 1 . We take f 2 : = max { 0 , z d , n e f } and e 2 : = min { e 1 , z d , n f 1 } , # S 5 = f 2 and # S 4 = e 2 and # S 3 = max { 0 , z d , n e 2 f 2 } .
(d) Assume z d , n < e + f . We take f : = max { 0 , z d , n e } and S 1 = .   □

5. Segre–Veronese Varieties

In this section, we fix integers k 2 and n i > 0 , i = 1 , , k . Set n : = n 1 + + n k and Y : = P n 1 × P n k . We have dim Y = n , and Y is the multiprojective space associated to tensors of format ( n 1 + 1 ) × × ( n k + 1 ) . Write G : = Aut ( P n 1 ) × × Aut ( P n k ) . We have G Aut ( Y ) and g * ( O Y ( x 1 , , x k ) ) O Y ( x 1 , , x k ) for all ( x 1 , , x k ) Z k .
Remark 10.
We discussed our approach in Remark 6. For Segre–Veronese varieties, the reduction is from an arbitrary A S ( Y , a ) to a specific B S ( Y , a ) contained in a curve L of multidegree ϵ 1 (Lemma 6). This is a key step because using the Differential Horace Lemma twice with respect to hypersurfaces containing L, we pass from A or B to ∅. To avoid several very dense pages, we reduce to results known in the literature, (ref. [] Cor. 3.14) for this section, ref. [] for the next one. The interested reader may use the published literature in a similar way in the following very interesting settings:
1. 
k = 2 , arbitrary biprojective space, bidegree ( d 1 , d 2 ) with d 1 3 and d 3 3 []. A few cases with d 1 = 2 and d 2 5 are in [].
2. 
k 3 , arbitrary multiprojective space, multidegree ( d 1 , , d k ) with d 1 3 , d 2 3 , d i 2 for all i > 2 [,]. A few cases with d k = 1 are in [].
3. 
k = 2 , d 1 = 1 . This part is wide open even for A = [,,].
Lemma 6.
Take A S ( Y , a ) such that # π 1 ( A ) = a . Fix p i P n i , 2 i k , and a line R P n 1 . Set L : = R × p 2 × × p k . Then there is B L such that # B = a , and B is a flat limit of set G-equivalents to A.
Proof. 
The case k = 1 (omitted in the statement) is true by the proof of Lemma 3. Hence, it is sufficient to find an irreducible family of elements of G that for i = 2 , , k sends the i-th coordinate of all points of A to a family with p i as its limit, preserving the condition that the first coordinate of all points of A are distinct. For i = 2 , , k , take a general hyperplane H i P n i . Since H i is general, we have p i H i . For i = 2 , , k , let Y ( H i ) denote the multiprojective space with H i as its i-th factor and P n j as its j-th factor for all j { 1 , , k } { i } . Set Y ˜ : = i = 2 k Y ( H i ) . Let l : Y Y ˜ P n 1 × p 2 × × p k be defined by the formula l ( ( x 1 , , x k ) ) = ( x 1 , p 1 , , p k ) . Since each H i is general, A Y ˜ = , and hence, is defined at each point of A. Since # π 1 ( A ) = a , we have # l ( A ) = a . Write l = ( l 1 , , l k ) with l 1 : P n 1 P n 1 the identity map and l i : P n i H i p i the linear projection from H i for i = 2 , , k . Fix i { 2 , , k } . In the proof of Lemma 3 (taking a system of homogeneous coordinates of P n i with p i a coordinate points), we proved that l i is a flat limit of a family { g ( i ) c } c K { 0 } of automorphisms of P n i . Write g ( 1 ) c : = l 1 for all c K { 0 } . Note that G = Aut ( P n 1 ) × × Aut ( P n k ) . Thus, for all c K { 0 } the automorphism g c : = ( g ( 1 ) c , , g ( k ) c ) is an element of G. By construction, the set l ( A ) is the flat limit of the family { g c ( A ) } c K { 0 } . If n 1 = 1 , the proof is over. If n 1 > 1 , we apply Lemma 3 to the first factor of Y.   □
Lemma 7.
Fix positive integers d 1 , , d k , a and x. Take A S ( Y , a ) such that # π 1 ( A ) = a . Fix p i P n i , 2 i k and a line R P n 1 . Set L : = R × p 2 × × p k . Then there is B S ( L , a ) such that for a general S S ( Y , x ) , we have h 0 ( I 2 B 2 S ( d 1 , , d k ) ) h 0 ( I 2 A 2 S ( d 1 , , d k ) ) and h 0 ( I B 2 S ( d 1 , , d k ) ) h 0 ( I A 2 S ( d 1 , , d k ) ) .
Proof. 
Use the semicontinuity theorem for cohomology (ref. [], Th. III.12.8) and Lemma 6.   □
Remark 11.
Take a = 1 and a positive integer x. Fix k, n i and d i , 1 i k . Take a general element ( S , S ) of the set S ( Y , x ) × S ( Y , x + 1 ) . Since G is transitive, we have the equality h 0 ( I 2 A 2 S ( d 1 , , d k ) ) = h 0 ( I 2 S ( d 1 , , d k ) ) and the equality
h 0 ( I A 2 S ( d 1 , , d k ) ) = max { 0 , h 0 ( I 2 S ( d 1 , , d k ) ) 1 , 0 } .
Hence, it is important to know when a secant variety of a Segre–Veronese variety is defective. Many cases are known [,,,]. We list the following 2 cases that will be needed in this paper. We call h 0 the integer h 0 ( I 2 S ( d 1 , , d k ) ) for a general S S ( Y , x ) and
δ : = deg ( 2 S ) h 0 ( O Y ( d 1 , , d k ) ) + h 0 = h 0 ( I 2 S ( d 1 , , d k ) ) .
(a) Assume n i = 1 for all i. With no loss of generality, we may assume d 1 d k . By [], the 4 exceptional cases are as follows:
1. 
k = 2 , d 1 = 2 , d 2 = 2 b with b a positive integer, x = 2 b + 1 , h 0 = 1 , δ = 1 ;
2. 
k = 3 , d 1 = d 2 = 1 , d 3 = 2 b with b a positive integer, x = 2 b + 1 , h 0 = 1 , δ = 1 ;
3. 
k = 3 , d 1 = d 2 = d 3 = 2 , x = 7 , h 0 = 1 , δ = 2 ;
4. 
k = 4 , d 1 = d 2 = d 3 = d 4 = 1 , x = 3 , h 0 = 2 , δ = 1 .
(b) Assume k = 2 , n 2 = 1 . If d 2 3 , then the associated Segre–Veronese variety is defective if and only if d 1 = 2 and d 2 is even (ref. [], Cor. 3.14).
Remark 12.
Take a = 2 . The group G acts transitively on the subset of S ( Y , 2 ) formed by 2 points not contained in a curve of multidegree ϵ 1 . Hence, for A S ( Y , 2 ) not contained in a curve of multidegree ϵ 1 , we have h 0 ( I A 2 S ( d 1 , , d k ) ) = max { 0 , h 0 ( I 2 S ( d 1 , , d k ) ) 2 } and h 0 ( I 2 A 2 S ( d 1 , , d k ) ) = h 0 ( I 2 S ( d 1 , , d k ) ) for a general S S ( Y , x ) and a general S S ( Y , x + 2 ) .
Remark 13.
The group G acts transitively on the set of all curves L Y of multidegree ϵ 1 . Fix any such curve L and call G L the set of all g G such that g ( L ) = L . The group G L acts 3-transitively on L. Thus, if a 3 and A , B S ( L , a ) , then h 0 ( I A 2 S ( d 1 , , d k ) ) = h 0 ( I B 2 S ( d 1 , , d k ) ) and h 0 ( I 2 A 2 S ( d 1 , , d k ) ) = h 0 ( I 2 B 2 S ( d 1 , , d k ) ) . Taking B general in S ( L , a ) , we obtain
h 0 ( I A 2 S ( d 1 , , d k ) ) = max { h 0 ( I 2 S ( d 1 , , d k ) ) a , h 0 ( I L 2 S ( d 1 , , d k ) ) } .
Remark 14.
Set Y : = P 1 × P 1 , i.e., take k = 2 and n 1 = n 2 = 1 . Fix positive integers a, d 1 , d 2 , x such that d 1 max { 1 , a 1 } . The triple ( d 1 , d 2 , 1 ) is not defective by Remark 11 and the fact that in case k = 2 of part (a) of Remark 11, all triples have defect at most 1. Now assume a 2 , d 1 = a 1 and A contained in a curve L of multidegree ϵ 1 . We have L P 1 and deg ( O L ( x 1 , x 2 ) ) = x 1 for all ( x 1 , x 2 ) Z 2 . For all integers x 1 , x 2 , all zero-dimensional schemes Z Y , we have the exact sequence
0 I Res L ( Z ) ( x 1 , x 2 1 ) I Z ( x 1 , x 2 ) I Z L , L ( x 1 , x 2 ) 0
We have h 0 ( I A ( d 1 , d 2 ) ) = h 0 ( O Y ( d 1 , d 2 1 ) ) by the long cohomology exact sequence associated to (6). Hence, the associated linear system | I A 2 S ( d 1 , d 2 ) | is defective for a general S S ( Y , x ) if and only if either d 2 = 1 and 2 x d 1 or d 2 = 3 , d 1 = a 1 is even and x = a or d 1 = 2 , a = 3 , d 2 is odd, d 2 3 and x = d 2 .
Remark 15.
Take an integral and non-degenerate variety X P r , an integral curve C X and integers a 0 , b 0 . Let Z C reg be a general union of a points and b double points of C reg . By Remark 5 applied to C C , we have dim Z = min { a + 2 b 1 , dim C } .
Lemma 8.
Fix integers a 1 , d 1 max { 2 a 2 , 2 } , d 2 2 and x > 0 . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , d 2 ) ) = 0 or h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 .
Proof. 
By Remark 7, it is sufficient to use the case in which A is contained in a curve L Y of multidegree ϵ 1 .
The case a = 1 is true by Remark 11 and the fact that in part (a), case k = 2 , of Remark 11, the defect is at most 1.
Thus, we may assume a 2 . Set
y : = ( d 1 + 1 ) ( d 2 + 1 ) / 3 a , ϵ : = ( d 1 + ) ( d 2 + 1 ) a 3 y .
Since ( d 1 + 1 ) ( d 2 + 1 ) 3 y 3 a 2 , to prove the lemma for a fixed triple ( d 1 , d 2 , A ) and all integers x, it is sufficient to prove that h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 for a general S S ( Y , y ) (Remarks 3 and 7).
Since a 2 , 2 a 2 a with equality if and only if a = 2 .
Assume a = 2 . By Remark 13, to prove the lemma, it is sufficient to prove the inequality h 0 ( I L 2 S ( d 1 , d 2 ) ) h 0 ( I 2 S ( d 1 , d 2 ) ) 2 , i.e., h 0 ( I 2 S ( d 1 , d 2 1 ) ) h 0 ( I 2 S ( d 1 , d 2 ) ) 2 , and that h 1 ( I 2 S ( d 1 , d 2 1 ) ) = 0 . Note that h 0 ( O Y ( d 1 , d 2 ) ) h 0 ( O Y ( d 1 , d 2 1 ) ) = d 1 + 1 3 . First, assume d 2 3 . Since d 2 3 , O Y ( d 1 , d 2 ) is not defective if d 1 3 , and hence, we conclude by the case k = 2 of Remark 11. Now, assume d 1 = 2 . Note that we have y = d 2 + 1 a / 3 = d 2 and ϵ = 1 . We use that in the exceptional cases with k = 2 of Remark 11, we have δ = 1 . Now, assume d 2 = 2 . Since a = 2 , we have y = d 1 and ϵ = 1 . We have h 0 ( I 2 S ( d 1 , 2 ) ) = 3 and h 0 ( I 2 S ( d 1 , 1 ) ) = 0 .
Assume a > 2 . Set e : = ( d 1 + 1 a ) / 2 and f : = d 1 + 1 a 2 e . We have 0 f 1 .
Assume for the moment y e + f .
Take a general ( S 1 , S 2 , S 3 ) S ( Y , y e f ) × S ( L , e ) × S ( L , f ) with the convention S 1 = if y e f = 0 , S 2 = if e = 0 and S 3 = if f = 0 . Note that we have deg ( ( 2 S 2 , L ) S 3 A ) = d 1 + 1 . By the Differential Horace Lemma (Lemma 2), it is sufficient to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , L ) ( d 1 , d 2 1 ) ) = 0 .
Observation 1: Since a + 2 e + f = d 1 + 1 and 3 y + a + ϵ = ( d 1 + 1 ) ( d 2 + 1 ) , we have deg ( 2 S 1 S 2 ( 2 S 3 , L ) ) = d 2 ( d 1 + 1 ) ϵ .
(a) In this step, we prove that h 1 ( I 2 S 1 ( d 1 , d 2 1 ) ) = 0 . We have 0 ϵ 2 and f 1 . We have deg ( 2 S 1 ) = 3 y 3 e 3 f = ( d 1 + 1 ) ( d 2 + 1 ) a ϵ 3 e 3 f and a + 2 e + f = d 1 + 1 . Hence, 3 y d 1 ( d 2 + 1 ) if and only if either e + ϵ f or e = 0 , f = 1 and ϵ > 0 . We have e = 0 and f = 1 if and only if a = d 1 , which is false because a 3 and d 1 2 a 2 . Hence, h 1 ( I 2 S 1 ( d 1 , d 2 1 ) ) = 0 if ( d 1 , d 2 1 , y ) is not in the list of Remark 11, part (a) for k = 2 . Since d 1 4 , the triple ( d 1 , d 2 1 , y ) may be in the list only if d 2 1 = 2 and d 1 is even, and in that case, the defect is at most 1. To conclude, it would be sufficient to have ϵ + e f > 0 . For d 1 = 4 , we have a = 3 , and hence, e = 1 and f = 0 . For d 1 = 6 , we have a = 4 , e = f = 1 , y = 6 and ϵ = 2 . For d 1 8 , we have e 2 > f .
(b) By Lemma 1, Remark 15 and Observation 1 to conclude for y e + f , it is sufficient to prove that h 0 ( I 2 S 1 ( d 1 , d 2 2 ) ) max { 0 , d 2 ( d 1 + 1 ) 3 y e 2 f } .
First, assume d 2 = 2 . We have y = d 1 + 1 a / 3 and ϵ = 3 a / 3 a . Since S 1 is general, h 0 ( I 2 S 1 ( d 1 , 0 ) ) = max { 0 , d 1 + 1 2 y } , we have d 1 + 1 2 y 0 , because 2 y = 2 d 1 + 2 2 a / 2 , d 1 2 a 2 and a 3 .
Now, assume d 2 3 and that ( d 1 , d 2 2 ) is not defective. We obtain
h 0 ( I 2 S 1 ( d 1 , d 2 2 ) ) = max { 0 , ( d 1 + 1 ) ( d 2 1 ) 3 y } ,
concluding this case.
Now, assume d 2 3 and that ( d 1 , d 2 2 , y ) is defective. By Remark 11 and the inequality d 1 3 , we have d 2 = 4 , d 1 4 , and d 1 is even. Moreover, the defect is 1 only if y = d 1 + 1 . We have 3 y + 2 a + ϵ = 5 d 1 + 5 with 0 ϵ 2 . We obtain 3 + 2 a + ϵ = 5 d 1 + 5 with d 1 2 a 2 and a 3 , a contradiction.
(c) Now, assume y e + f 1 . We take S 3 = , S 2 of cardinality min { e , y } and S 1 of cardinality y min { e , y } .   □
Proposition 2.
Set Y : = P 1 × P 1 , i.e., take k = 2 and n 1 = n 2 = 1 . Fix positive integers a, d 1 , d 2 , x such that d 1 max { 3 , 2 a 1 } and d 2 3 . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , d 2 ) ) = 0 or h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 .
Proof. 
By Remark 7, it is sufficient to use the case in which A is contained in a curve L Y of multidegree ϵ 1 .
The case a = 1 is true by Remark 11 and in particular (part (a), case k = 2 of Remark 11) for d 1 3 and d 2 3 , no case is defective. Thus, we may assume a 2 . Set y : = ( d 1 + 1 ) ( d 2 + 1 ) / 3 a and ϵ : = ( d 1 + 1 ) ( d 2 + 1 ) 3 a 3 y . Since ϵ 2 , to prove the lemma for a fixed triple ( d 1 , d 2 , A ) and all integers x, it is sufficient to prove that h 1 ( I 2 A 2 S ( d 1 , d 2 ) ) = 0 for a general S S ( Y , y ) (Remarks 3 and 4).
First, assume d 1 = 2 a 1 . Since deg ( L 2 A ) = 2 a , deg ( O L ( d 1 , d 2 ) ) = d 1 and L P 1 , the long cohomology exact sequence of (6) gives h i ( I 2 A 2 S ( d 1 , d 2 ) ) = h i ( I A 2 S ( d 1 , d 2 1 ) ) . Hence, Lemma 8 concludes this case. Thus, we may assume d 1 2 a .
Set e : = ( d 1 + 1 2 a ) / 2 and f : = d 1 + 1 2 a 2 e . We have 0 f 1 . Until step (c), we assume y e + f . Take a general ( S 1 , S 2 , S 3 ) S ( Y , y e f ) × S ( L , e ) × S ( L , f ) with the convention S 1 = if y e f = 0 , S 2 = if e = 0 and S 3 = if f = 0 . We have deg ( ( 2 S 2 , L ) S 3 ( 2 A , L ) ) = d 1 + 1 . By the Differential Horace Lemma (Lemma 2), it is sufficient to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , L ) ( d 1 , d 2 1 ) ) = 0 .
Observation 1: Since 2 a + 2 e + f = d 1 + 1 and 3 y + 3 a + ϵ = ( d 1 + 1 ) ( d 2 + 1 ) , we have deg ( 2 S 1 S 2 ( 2 S 3 , L ) A ) = d 2 ( d 1 + 1 ) ϵ .
(a) In this step, we prove that h 1 ( I 2 S 1 A ( d 1 , d 2 1 ) ) = 0 . Since d 2 1 2 , Lemma 8 gives that it is sufficient to prove that 3 ( y e f ) + a ( d 1 + 1 ) d 2 . Observation 1 gives 3 ( y e f ) + a = d 2 ( d 1 + 1 ) ϵ e .
(b) In this step, we conclude the case y e + f . By step (a), we have the equality h 0 ( I 2 S 1 A ( d 1 , d 2 1 ) ) = d 2 ( d 1 + 1 ) a 3 ( y e f ) . By step (a), Observation 1 and Lemma 1, it is sufficient to prove that
h 0 ( I 2 S 1 ( d 1 , d 2 2 ) ) max { 0 , d 2 ( d 1 + 1 ) a 3 y + 2 e + f } .
Since S 1 is general, this inequality is obvious if ( d 1 , d 2 2 , y e f ) is not defective. Since d 1 3 and d 2 3 , ( d 1 , d 2 2 , x ) is defective if and only if d 2 = 4 , d 1 is even and x = d 1 + 1 , and in that case the defect is 1 and occurs only if h 0 ( I 2 S ( d 1 , d 2 ) ) = 1 (Remark 11). Assume d 2 = 4 . Since h 0 ( O Y ( d 1 , 3 ) ) h 0 ( O Y ( d 1 , 1 ) ) , it is sufficient to prove that a + e + 2 f d 1 . This inequality is obvious, because f 1 , a 3 and 2 a + 2 e + f = d 1 + 1 .
(c) Assume y e + f 1 . We take S 3 = , S 2 of cardinality min { y , e } and S 1 of cardinality y max { e , y } .   □
Remark 16.
Theorem 5 and Proposition 3 are not optimal because we made the assumption d 1 5 to use Theorem 2. To use the case with d 1 { 3 , 4 } , one needs to first obtain all exceptional cases for d = 3 , 4 for the Veronese variety. By [], Table 1 at p. 91, we need d 1 3 and d 2 3 to avoid all exceptional cases for the case a 1 .
Theorem 5.
Fix integers a 0 , d 1 max { 4 , 2 a 1 } , d 2 3 and m 1 . Set Y : = P m × P 1 . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , d 2 ) ) = 0 or h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 .
Lemma 9.
Fix integers a 0 , d 1 3 + a , d 2 3 and x > 0 . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , d 2 ) ) = 0 or h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 .
Proof. 
By Remark 7, it is sufficient to use the case in which A is contained in a curve L Y of multidegree ϵ 1 .
The case a is true by Remark 11 and the fact that in part (a), case k = 2 , because d 1 3 and d 2 3 . Thus, we may assume a 2 . Set y : = ( d 1 + 1 ) ( d 2 + 1 ) / 3 a and ϵ : = ( d 1 + ) ( d 2 + 1 ) a 3 y . Since ( d 1 + 1 ) ( d 2 + 1 ) 3 y 3 a 2 , to prove the lemma for a fixed triple ( d 1 , d 2 , A ) and all integers x, it is sufficient to prove that h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 for a general S S ( Y , y ) (Remark 7).
Assume a = 2 . By Remark 13, it is sufficient to prove two statements, the inequality h 0 ( I L 2 S ( d 1 , d 2 ) ) h 0 ( I 2 S ( d 1 , d 2 ) ) 2 , i.e., h 0 ( I 2 S ( d 1 , d 2 1 ) ) h 0 ( I 2 S ( d 1 , d 2 ) ) 2 , and that h 1 ( I 2 S ( d 1 , d 2 1 ) ) = 0 . Note that h 0 ( O Y ( d 1 , d 2 ) ) h 0 ( O Y ( d 1 , d 2 1 ) ) = d 1 + 1 3 . First, assume d 2 3 . Since d 2 3 , O Y ( d 1 , d 2 ) is not defective by the inequality d 1 3 , hence, we conclude by the case k = 2 of Remark 11. Assume a > 2 . Set e : = ( d 1 + 1 a ) / 2 and f : = d 1 + 1 a 2 e . We have 0 f 1 .
Assume for the moment y e + f .
Take a general ( S 1 , S 2 , S 3 ) S ( Y , y e f ) × S ( L , e ) × S ( L , f ) with the convention S 1 = if y e f = 0 , S 2 = if e = 0 and S 3 = if f = 0 .
We have deg ( ( 2 S 2 , L ) S 3 A ) = d 1 + 1 . By the Differential Horace Lemma (Lemma 2), it is sufficient to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , L ) ( d 1 , d 2 1 ) ) = 0 .
Observation 1: Since a + 2 e + f = d 1 + 1 and 3 y + a + ϵ = ( d 1 + 1 ) ( d 2 + 1 ) , we have deg ( 2 S 1 S 2 ( 2 S 3 , L ) ) = d 2 ( d 1 + 1 ) ϵ .
(a) In this step, we prove that h 1 ( I 2 S 1 ( d 1 , d 2 1 ) ) = 0 . The only difference with respect to the proof of Lemma 6 is that to exclude that ( e , f ) = ( 0 , 1 ) , it is sufficient to use that a + 2 e + f = d 1 + 1 with d 1 3 .
(b) By Lemma 1, Remark 15 and Observation 1, to conclude for y e + f , it is sufficient to prove that h 0 ( I 2 S 1 ( d 1 , d 2 2 ) ) max { 0 , d 2 ( d 1 + 1 ) 3 y e 2 f } . We have d 2 3 + a 5 , excluding almost all cases. The remaining ones are trivial by Remark 5.
(c) Now assume y e + f 1 . We take S 3 = , S 2 of cardinality min { e , y } and S 1 of cardinality y min { e , y } .   □
Proposition 3.
Fix a positive integer m and set Y : = P m × P 1 . Take integers a 0 , x, d 1 3 + a and d 2 3 . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , d 2 ) ) = 0 or h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 .
Proof. 
Since the case m = 1 is true by Lemma 9, we assume m 2 and use induction on m. By Remark 7, it is sufficient to use the case in which A is contained in a curve L Y of multidegree ϵ 1 . We have dim Y = m + 1 and h 0 ( O Y ( d 1 , d 2 ) ) = m + d 1 m ( d 2 + 1 ) .
Set y : = ( ( m + d 1 m ) ( d 2 + 1 ) a ) / ( m + 2 ) , x 1 : = ( ( m + d 1 m ) ( d 2 + 1 ) a ) / ( m + 2 ) and ϵ m + d 1 m ( d 2 + 1 ) a ( m + 2 ) y . By Remark 2, to prove the proposition for a fixed ( Y , d 1 , d 2 , A ) and all x, it is sufficient to prove it for x = y and x = x 1 . In steps (a) and (b), we use the case z = y . In step (c), we handle the case x = x 1 . Take a general S S ( Y , y ) .
By Remark 11 and [], Cor 3.14, the case a = 1 is true. Hence, we may assume a 2 . Since A L , there is M | I A ( 1 , 0 ) | . We have M P m 1 × P 1 and a m + d 1 1 m 1 ( d 2 + 1 ) . Set e : = ( ( m + d 1 1 m 1 ) ( d 2 + 1 ) a ) / ( m + 1 ) and f : = m + d 1 1 m 1 ( d 2 + 1 ) ( m + 1 ) e a . We have 0 f m .
Observation 1: Assume a = 2 . If # π 2 ( A ) , then the set A is in the open orbit of S ( Y , 2 ) for the action of the group G considered in Remark 11. Hence, this case follows for the non-defectivity of Y for the linear system | O Y ( t 1 , t 2 ) for all t 1 3 , t 2 3 (ref. [], Cor. 3.14).
(a) Assume y e + f .
Take a general ( S 1 , S 2 , S 3 ) S ( Y , y e f ) × S ( M , e ) × S ( M , f ) . Note that deg ( A ( 2 S 2 , M ) S 3 ) = m + d 1 1 m 1 ( d 2 + 1 ) . By the inductive assumption on m, we have h i ( M , I A ( 2 S 2 , M ) S 3 , M ( d 1 , d 2 ) ) = 0 , i = 0 , 1 . We specialize e of the points of S to different points of S 2 and apply the Differential Horace Lemma to each point of S. We find that to prove that h 1 ( I A 2 S ( d 1 , d 2 ) ) = 0 , it is sufficient to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = 0 .
(a1) In this step, we prove that h 1 ( I A 2 S 1 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = 0 . By Remark 2, it is sufficient to prove that h 1 ( I 2 S 1 2 S 3 ( d 1 1 , d 2 ) ) = 0 . Note that here, A does not appear, except that M is a hyperplane containing L. We take as M a general hyperplane containing L. The group Aut ( M ) acts transitively on the set of of all lines of M. Hence, S 3 may be considered as a general subset of f points of M. Taking S 1 general, we find that 2 S 1 2 S 3 has the bigraded Hilbert function of the union of y e general double points of Y. By [], Cor. 3.14, to prove that h 1 ( I 2 S 1 2 S 3 ( d 1 1 , d 2 ) ) = 0 , it is sufficient to prove that ( m + 2 ) ( y e ) m + d 1 1 m 1 . Since deg ( A ( 2 S 2 , M ) S 3 ) = m + d 1 1 m 1 ( d 2 + 1 ) and ( m + 2 ) y + a = m + d 1 m ϵ , we have ( m + 2 ) ( y e ) = f ϵ e + m + d 1 1 m 1 ( d 2 + 1 ) . Hence, it is sufficient to prove that e f . Assume e f 1 . By the definition of e as a floor, we obtain m + d 1 1 m 1 1 + m + ( f 1 ) m . Since f m , we obtain m + d 1 1 m 1 1 + m 2 . It is sufficient to use that d 1 3 + a 5 .
(a2) Step (a1) gives
h 0 ( I 2 S 1 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = ( d 2 + 1 ) m + d 1 1 m ( m + 2 ) ( y e f ) ( m + 1 ) f .
By Lemma 2, to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = 0 , it is sufficient to prove that
h 1 ( I 2 S 1 ( d 1 2 , d 2 ) ) ( d 2 + 1 ) m + d 1 1 m ( m + 2 ) ( y e f ) ( m + 1 ) f e .
Since deg ( 2 S 1 S 2 ( 2 S 3 , M ) ) = h 0 ( O Y ( d 1 1 , d 2 ) ) ϵ with ϵ 0 , to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = 0 , it is sufficient to prove that
h 0 ( I 2 S 1 ( d 1 2 , d 2 ) ) max { 0 , d 2 + 1 ) m + d 1 1 m ( m + 2 ) ( y e f ) ( m + 1 ) f e } .
Since d 1 2 3 , d 2 3 and S 1 is general, we may apply [], Cor. 3.14, and find that either h 0 ( I 2 S 1 ( d 1 2 , d 2 ) ) = 0 or h 1 ( I 2 S 1 ( d 1 2 , d 2 ) ) = 0 .
Hence, we may assume that h 1 ( I 2 S 1 ( d 1 2 , d 2 ) ) = 0 , i.e., we may assume that h 0 ( I 2 S 1 ( d 1 2 , d 2 ) ) = h 0 ( O Y ( d 1 2 , d 2 ) ) 2 S 1 .
Note that h 0 ( O Y ( d 1 1 , d 2 ) ) h 0 ( O Y ( d 1 2 , d 2 ) ) = m + d 1 2 m 1 ( d 2 + 1 ) . Thus, it is sufficient to check that e f . Assume e f 1 . We find a + f 2 m + d 1 1 m 1 ( d 2 + 1 ) with d 2 3 , d 1 3 + a and f m , a contradiction.
(b) Assume y < e + f . We modify step (a), taking instead of the triple ( y e f , e , f ) , the following triple ( y , e , f ) with y = 0 , f : = max { 0 , f ( e + f y ) } and e : = y f .
(c) In this step, we handle the case x = x 1 . We may assume x 1 y , i.e., x 1 = y + 1 . We take the same e, f, S 2 and S 3 , but here, S 1 has cardinality x 1 e f . Assume for the moment x 1 e + f . In step (a1), the only modification arises if h 1 ( I 2 S 1 2 S 3 ( d 1 1 , d 2 ) ) > 0 . Assume h 1 ( I 2 S 1 2 S 3 ( d 1 1 , d 2 ) ) > 0 . By [], Cor. 3.14, we obtain h 0 ( I 2 S 1 2 S 3 ( d 1 1 , d 2 ) ) = 0 . Since S 2 is general in M, it is sufficient to use that e f .
If x 1 < e + f , the proof of step (b) works verbatim.   □
Proof of Theorem 5.
The cases with a 1 are true by [], Cor. 3.14, and Remark 11. Thus, we may assume a 2 , and hence, d 1 6 . The case m = 1 is true by Proposition 3. Thus, we assume m 2 and use induction on m.   □
We have dim Y = m + 1 and h 0 ( O Y ( d 1 , d 2 ) ) = m + d 1 m ( d 2 + 1 ) .
Set y : = ( m + d 1 m ) ( d 2 + 1 ) / ( m + 2 ) a , x 1 : = ( m + d 1 m ) ( d 2 + 1 ) / ( m + 2 ) a and ϵ m + d 1 m ( d 2 + 1 ) ) ( m + 2 ) ( y + a ) . By Remark 2, to prove the proposition for a fixed ( Y , d 1 , d 2 , A ) and all x, it is sufficient to prove it for x = y and x = x 1 . In steps (a) and (b), we use the case x = y . In step (c), we use the case x = x 1 . Take a general S S ( Y , y ) .
By Remark 7, it is sufficient to use the case in which A is contained in a curve L Y of multidegree ϵ 1 . Take a general M | I L ( 1 , 0 ) | . We have dim M = m and h 0 ( M , O M ( d 1 , d 2 ) ) = m + d 1 1 m 1 ( d 2 + 1 ) . Set e : = a + ( m + d 1 1 m 1 ) ( d 2 + 1 ) / ( m + 1 ) and f : = m + d 1 1 m 1 ( d 2 + 1 ) ( m + 1 ) ( e + a ) . We have 0 f m and
2 a + ( m + 1 ) e + f = ( d 2 + 1 ) m + d 1 1 m 1
(a) Assume y e + f .
Take a general ( S 1 , S 2 , S 3 ) S ( Y , y e f ) × S ( M , e ) × S ( M , f ) . Note that deg ( 2 A ( 2 S 2 , M ) S 3 ) = m + d 1 1 m 1 ( d 2 + 1 ) . By the inductive assumption on m, we have h i ( M , I 2 A ( 2 S 2 , M ) S 3 , M ( d 1 , d 2 ) ) = 0 , i = 0 , 1 . We specialize e of the points of S to different points of S 2 and apply the Differential Horace Lemma (Lemma 2) to each point of S. We find that to prove that h 1 ( I 2 A 2 S ( d 1 , d 2 ) ) = 0 , it is sufficient to prove that h 1 ( I 2 S 1 A S 2 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = 0 .
Set g : = ( ( m + d 1 2 m 1 ) a e ( m + 1 ) f ) / ( m + 1 ) and
h : = m + d 1 + m 2 m 1 ( d 2 + 1 ) a m f ( m + 1 ) g .
We have 0 h m and
a + e + ( m + 1 ) f + ( m + 1 ) g + h = ( d 2 + 1 ) m + d 1 2 m 1
Until step (a3), we assume y e + f + g + h . Take a general
( S 4 , S 5 , S 6 ) S ( Y , y e f g h ) × S ( Y , g ) × S ( Y , h ) .
Note that deg ( ( 2 S 5 , M ) ( 2 S 3 , M ) A S 6 ) ) = h 0 ( M , O M ( d 1 1 , d 2 ) ) and that d 1 1 3 + a . Hence, Proposition 3 gives h i ( M , I ( 2 S 5 , M ) ( 2 S 3 , M ) A S 6 ( d 1 1 , d 2 ) ) = 0 , i = 0 , 1 . We specialize g of the points of S 1 to different points of S 5 and apply the Differential Horace Lemma to each point of S 6 . Since h i ( M , I ( 2 S 5 , M ) ( 2 S 3 , M ) A S 6 ( d 1 1 , d 2 ) ) = 0 , i = 0 , 1 , to prove that h 1 ( I 2 S 1 A S 2 ( 2 S 3 , M ) ( d 1 1 , d 2 ) ) = 0 , it is sufficient to prove that
h 1 ( I 2 S 4 S 5 ( 2 S 6 , M ) ( d 1 2 , d 2 ) ) = 0 .
(a1) Assume y e + f + g + h . In this step, we prove the vanishing of the cohomology group H 1 ( I 2 S 4 ( 2 S 6 , M ) ( d 1 2 , d 2 ) ) = 0 . Note that here, A does not appear, except that M is a hyperplane containing L. We take as M a general hyperplane containing L. The group Aut ( M ) acts transitively on the set of of all lines of P m . Hence, S 3 may be considered as a general subset of f points of M. Taking S 1 general, we find that 2 S 4 2 S 6 has the bigraded Hilbert function of the union of y e f g general double points of Y. By [], Cor. 3.14, to prove that h 1 ( I 2 S 4 2 S 6 ( d 1 2 , d 2 ) ) = 0 , it is sufficient to prove that ( m + 2 ) ( y e f g ) m + d 1 1 m 1 .
Hence, h 1 ( I 2 S 4 S 5 ( 2 S 6 , M ) ( d 1 2 , d 2 ) ) = 0 .
(a2) In this step, we assume y e + f + g + h . In this step, we prove that h 1 ( I 2 S 4 S 5 ( 2 S 6 , M ) ( d 1 2 , d 2 ) ) = 0 . Note that deg ( S 5 ( 2 S 6 , M ) ) m + d 1 3 m 1 ( d 2 + 1 ) .
Observation 1: For all integers d 3 , we have
m + d 2 m 1 / m + d 3 m 1 = ( m + d 2 ) / ( d 1 ) .
Claim 4.
We have deg ( S 5 ( 2 S 6 , M ) ) m + d 1 3 m 1 ( d 2 + 1 ) .
Proof of Claim 4.
We have deg ( S 5 ( 2 S 6 , M ) ) = g + ( m + 1 ) h .
Assume g + ( m + 1 ) h 1 m + d 1 3 m 1 ( d 2 + 1 ) , and hence,
( m + 1 ) g + ( m + 1 ) 2 h m 1 ( m + 1 ) ( d 2 + 1 ) m + d 1 3 m 1 .
Set α : = m + d 1 3 m 1 . Multiplying (8) by (m + 1), subtracting (9) and using Observation 1, we obtain
( m + 1 ) a ( m + 1 ) e + ( m 2 + 2 m ) h m 1 α ( d 2 + 1 ) ( m + d 2 ) ( m + 1 ) d 1
Since h m , we obtain
( m + 1 ) a ( m + 1 ) e + m 3 + 2 m 2 m 1 α ( d 2 + 1 ) ( m + d 2 ) ( m + 1 ) d 1
To obtain a contradiction, it is sufficient to use that d 1 5 , d 2 3 and that (7) e m 2 / 3 .   □
Claim 1 and Lemma 1 show that to prove that h 1 ( I 2 S 4 S 5 ( 2 S 6 , M ) ( d 1 2 , d 2 ) ) = 0 , it is sufficient to prove that
h 0 ( I 2 S 4 ( d 1 3 , d 2 ) ) max { 0 , h 0 ( O Y ( d 1 2 , d 2 ) ) deg ( 2 S 4 S 5 ( 2 S 6 , M ) ) } .
This inequality is true because S 4 is general and O Y ( d 1 3 , d 2 ) is not defective (ref. [], Cor. 1.3).
(a3) Assume e + f y < e + f + g + h . We modify the proof of step (a2), taking ( g , h ) instead of ( g , h ) with h : = max { 0 , h ( e + f + g + h y ) } and # S 5 = g = y h .
(b) Assume y < e + f . We modify step (a1), taking instead of the previous sets S 1 , S 2 , S 3 the sets S 1 , S 2 , S 3 with S 1 = , # S 3 = max { 0 , f ( e + f y ) } and # S 2 = y # S 3 .
(c) Take x = x 1 . We may assume x 1 y , i.e., x 1 = y + 1 . We take the same e, f, g and h S 2 , S 3 , S 5 and S 6 , but here, # S 1 = x 1 e f and # S 6 = x 1 e f g h .

6. Segre–Veronese Varieties with One-Dimensional Factors

In this section, we prove the following theorem and several related results.
Theorem 6.
Set Y : = ( P 1 ) k , k 3 . Fix positive integers a, x, d i , 1 i k such that d 1 max { 5 , 2 a 1 } , d i 3 . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , , d k ) ) = 0 or h 1 ( I A 2 S ( d 1 , , d k ) ) = 0 .
To prove Theorem 6 (or similar statements) using a proof by induction, the following result may be very useful.
Theorem 7.
Set Y : = ( P 1 ) k , k 3 . Fix positive integers a, x and d i , 1 i k such that d 1 max { 3 , a 1 } , d 2 3 and d i 2 for all 3 i k . Take A S ( Y , a ) such that # π 1 ( A ) = a and a general S S ( Y , x ) . Then, either h 0 ( I A 2 S ( d 1 , , d k ) ) = 0 or h 1 ( I A 2 S ( d 1 , , d k ) ) = 0 .
To prove these theorems, we introduce the following notation and make the following remarks and results. Set d ̲ : = ( d 1 , , d k 1 ) N k 1 . We fix positive integers t 1 , , t k 1 such that t 1 a 1 and set t ̲ = ( t 1 , , t k 1 ) . In the proofs of Theorem 6 and 7, we take t i = d i for all i < k .
In Theorems 6 and 7, we assumed k 3 . We quote Proposition 3 and Theorem 5 for the case k = 2 and use induction on the integer k.
Remark 17.
By [] and the homogeneity of Y, we may assume a 2 . By Lemma 7, it is sufficient to use the case A L with L, a curve of multidegree ϵ 1 . Note that π k ( L ) is a point, o. Set D : = ( P 1 ) k 1 × { o } Y . Note that A L Y and that D | O Y ( ϵ k ) | .
Set α : = h 0 ( O D ( t 1 , , t k ) ) . We have α = i = 1 k 1 ( t i + 1 ) . Thus, for all t N , we have h 0 ( O Y ( t ̲ , t ) ) = α ( t + 1 ) . We have O D ( t ̲ , t ) O D ( t ̲ , 0 ) for all t Z . Remember that we will take t i : = d i in the proofs of Theorems 6 and 7, but we want to state and prove Theorems 8 and 9 in a more general way.
Let e, f be integers satisfying
a + k e + f = α ,
e + k f α and with α e k f minimum. Since k x + y = k ( x 1 ) + ( y + k ) and
( x 1 ) + k ( y + k ) = x + k y + k 2 1 , the integers e and f satisfy the following inequalities:
α k 2 + 1 e + k f α
If α a + k 2 1 , then ( e , f ) N 2 , since a > 0 , f e .
Remark 18.
Since t i 2 for all 2 i k 1 and t 1 max { 3 , a 1 } , we have α 2 k 2 ( t 1 + 1 ) and α a + k 2 .
Theorem 8.
Assume α a + k 2 , t 1 max { 3 , a 1 } and that ( D , t ̲ ) is not secant defective. Assume that for all b N and a general B S ( D , b ) , either h 0 ( D , I A 2 B , D ( t ̲ ) ) = 0 or h 1 ( D , I A 2 B , D ( t ̲ ) ) = 0 . Take a general S S ( Y , e + f ) and a general S S ( Y , e + f + k ) . Then, h 1 ( I A 2 S ( t ̲ , 1 ) ) = 0 , h 0 ( I A 2 S ( t ̲ , 1 ) ) = α k f e k 2 1 and h 0 ( I A 2 S ( t ̲ , 1 ) ) = 0 .
Proof. 
Since A L , t 1 max { 3 , a 1 } and t i > 0 for all i < k , we have h 1 ( D , I A , D ( t ̲ ) ) = 0 . Hence, the assumptions on A 2 B are not satisfied for the trivial reason that we have h 0 ( D , I A , D ( t ̲ ) ) = 0 .
By (12) and (13), to prove the result for S, it is sufficient to prove the equality h 1 ( I A 2 S ( t ̲ , 1 ) ) = 0 . Take a general ( S 1 , S 2 ) S ( D , e ) × S ( D , f ) . By the assumption on I A , D ( t ̲ ) and (12), we have the equalities h i ( D , I A 2 S 1 S 2 ( t ̲ , 1 ) ) = 0 , i = 0 , 1 , h 1 ( D , I S 1 2 S 2 ( d ̲ , 0 ) ) = 0 and h 0 ( D , I S 1 2 S 2 ( d ̲ , 1 ) ) = α e k f . We degenerate e of the points of S to different points of S 1 and apply the Differential Horace Lemma to each point of S 2 . We obtain h 1 ( I A 2 S ( t ̲ , 1 ) ) h 1 ( I S 1 2 S 2 ( t ̲ , 0 ) ) = 0 .
We degenerate S to ( S 1 , S 2 ) with S 2 general in S ( D , f + k ) and use the assumption that ( D , t ̲ ) is not secant defective to obtain h 0 ( D , I S 1 ( 2 S 2 , D ) , D ( t ̲ , 0 ) ) = 0 .   □
Remark 19.
Take the assumptions of Theorem 8 and add the conditions that α 0 mod k + 1 and a 0 ( mod k 2 1 ) . Set e ˜ : = a k + 1 k a k 2 1 and f ˜ : = α k + 1 + a k 2 1 . In the proof of Theorem 11, take a general ( S ˜ 1 , S ˜ 2 ) S ( D , e ˜ ) × S ( D , f ˜ ) instead of ( S 1 , S 2 ) . The same proof gives h i ( I S ( t ̲ , 1 ) ) = 0 , i = 0 , 1 , for a general S S ( Y , ( 2 α a ) / ( k + 1 ) ) .
Lemma 10.
Take the assumptions of Theorem 8 with α a + 2 ( k + 1 ) 2 . We have e + f k . Fix an integer z such that 0 z k . Let Z Y be a general union of e + f z double points of Y and z double points of D. Then, h 1 ( I A Z ( t ̲ , 1 ) ) = 0 .
Proof. 
We have ( k + 1 ) ( e + f ) 2 α a k 2 + 1 . We follow the proof of Theorem 8 taking instead of 2 S 1 2 S 2 , the same reduction, S 1 S 2 , but with z of its connected component general double points of D, not general double points of Y with reduction general in D.   □
Proof of Theorem 7.
Since the case k = 2 is true by Proposition 3, we assume k > 2 and use induction on the integer k. Fix D | O Y ( ϵ k ) | . We take t i = d i for i < k in the statement of Theorem 8. Hence, α = i = 1 k 1 ( d i + 1 ) . By [], ( D , d ̲ ) is not secant defective. Since a 2 , d 1 a 1 and d 2 3 , we have α a + 2 ( k + 1 ) 2 , and hence, we may apply Theorem 8 and Lemma 10.
(a) In this step, we use the case d k = 2 . Let w be the maximal integer such that h 1 ( I A 2 S ( d ̲ , 1 ) ) = 0 and set β : = 2 α a ( k + 1 ) w . By Theorem 8, we have w e + f and β k 2 1 . Set y : = 3 α / ( k + 1 ) and x 1 : = 3 α / ( k + 1 ) . By Remark 2, it is sufficient to prove that h 1 ( I A 2 S ( d ̲ , 2 ) ) = 0 for a general S S ( Y , y ) and h 0 ( I A 2 S ( d ̲ , 2 ) ) = 0 for a general S S ( Y , x 1 ) .
(a1) In this step, we prove that h 1 ( I A 2 S ( d ̲ , 2 ) ) = 0 for a general S S ( Y , y ) . Take a general S S ( Y , y ) . Set g : = ( α a ) / k and h : = α a k g . We have 0 h k 1 .
(a1.1) Assume y g + h . Take a general
( S 1 , S 2 , S 3 ) S ( Y , y g h ) × S ( D , g ) × S ( D , h ) .
By the assumptions on D and d ̲ , we have h i ( D , I ( 2 S 2 , D ) S 3 A , D ( d ̲ ) ) = 0 , i = 0 , 1 . We degenerate g of the points of S to different points of S 2 and apply the Differential Horace Lemma to each point of S 3 . By Lemma 2, to prove that h 1 ( I A 2 S ( d ̲ , 2 ) ) = 0 , it is sufficient to prove that h 1 ( I 2 S 1 S 2 ( 2 S 3 , D ) ( d ̲ , 1 ) ) = 0 .
We prove that h 1 ( I 2 S 1 ( 2 S 3 , D ) ( d ̲ , 1 ) ) = 0 . Since # S 3 k , to apply Lemma 10, it is sufficient to prove that y g e + f . To handle the case with x = x 1 , we prove the following stronger numerical result.   □
Claim 5.
We have x 1 e + f + g .
Proof of Claim 5.
Assume x 1 g + e + f + 1 . Since ( k + 1 ) x 1 3 α + k , we find 3 k α k ( k + 1 ) g + k ( k + 1 ) e + k ( k + 1 ) f + 1 . Since k e + f = α a and k g + h = α , we find 3 k α ( 2 k + 2 ) α 2 ( k + 1 ) a + ( k 2 1 ) f ( k + 1 ) h . Since h k , we find a contradiction.   □
Claim 1 and Lemma 10 with z : = h give h 1 ( I 2 S 1 ( 2 S 3 , D ) ( d ̲ , 1 ) ) = 0 , and hence, h 0 ( I 2 S 1 ( 2 S 3 , D ) ( d ̲ , 1 ) ) = 2 α ( k + 1 ) . By Lemma 1, to prove that h 1 ( I 2 S 1 ( 2 S 3 , D ) ( d ̲ , 1 ) ) = 0 , it is sufficient to prove that h 0 ( I 2 S 1 ( d ̲ , 0 ) 2 α ( k + 1 ) g . Since S 1 is general in Y, its image by the projection of Y onto its first factors is general. Since ( ( P 1 ) k 1 , d ̲ ) is not secant defective, h 0 ( I 2 S 1 ( d ̲ , 0 ) ) = max { 0 , α k ( y k g ) } . By Lemma 1, to conclude for y, it is sufficient to prove that 2 α ( k + 1 ) ( y g h ) . Since ( k + 1 ) y 3 α , it is sufficient to prove that ( k + 1 ) ( g + h ) α . This inequality is true because k g + h = α a .
(a1.1) Assume y < g + h . Instead of h, take h : = max { 0 , g + h y } , g : = y h and S 1 = .
(a2) The case x = x 1 is easier, because we proved Claim 1 for it and we only need to check that h 0 ( I 2 S 1 ( d ̲ , 0 ) ) max { 0 , α k ( y k g ) } .
(b) In this step, we take d k > 2 and prove this case by induction on the integer d k . To prove the inductive step, we use the case d k 1 of Lemma 10. Set y : = d k α / ( k + 1 ) . For all integers z such that 0 z k , we check that h 1 ( I A Z ( t ̲ , d k 1 ) ) = 0 with Z a general union of y z double points of Y and of z double points of D.
Let e 1 , f 1 be integers satisfying
k a + k e 1 + f 1 = α ,
a + e 1 + k f 1 α such that α a e 1 k f 1 is minimal. We saw that
α k 2 + 1 a + e 1 + k f 1 α
We have f 1 e 1 . If α k a + k 2 1 , then e 1 0 .
Theorem 9.
Assume α k a + k 2 , t 1 max { 5 , 2 a 1 } and that ( D , t ̲ ) is not secant defective. Assume that for all b N and a general B S ( D , b ) , either h 0 ( I 2 A 2 B , D ( t ̲ ) ) = 0 or h 0 ( I 2 A 2 B , D ( t ̲ ) ) = 0 . Take a general S S ( Y , e 1 + f 1 ) and a general S S ( Y , e 1 + f 1 + k ) . Then, h 1 ( I 2 A 2 S ( t ̲ , 1 ) ) = 0 , h 0 ( I 2 A 2 S ( t ̲ , 1 ) ) = α k f e k 2 1 and h 0 ( I 2 A 2 S ( t ̲ , 1 ) ) = 0 .
Proof. 
Since A L , t 1 2 a 1 and t i > 0 for all i < k , we have h 1 ( D , I 2 A , D ( t ̲ ) = 0 . Hence, the assumptions on 2 A 2 B are not satisfied for the trivial reason that h 0 ( I 2 A ( t ̲ ) ) = 0 .
By (14) and 15, to prove the result for S, it is sufficient to prove that h 1 ( I 2 A 2 S ( t ̲ , 1 ) ) = 0 . Take a general ( S 1 , S 2 ) S ( D , e 1 ) × S ( D , f ) 1 . By the inductive assumption on k and (14) and (15), we have the equalities h i ( D , I 2 A 2 S 1 S 2 ( t ̲ , 1 ) ) = 0 , i = 0 , 1 , h 1 ( D , I A S 1 2 S 2 ( t ̲ , 0 ) ) = 0 and h 0 ( D , I S 1 2 S 2 ( t ̲ , 0 ) ) = α e 1 k f . We degenerate e 1 of the points of S to different points of S 1 and apply the Differential Horace Lemma to each point of S 2 .
We obtain h 1 ( I 2 A 2 S ( t ̲ , 1 ) ) h 1 ( I A S 1 2 S 2 ( t ̲ , 0 ) ) = 0 .
We degenerate S to ( S 1 , S 2 ) with S 2 general in S ( D , f + k ) and use the inductive assumption to obtain h 0 ( D , I S 1 ( 2 S 2 ) , D ( t ̲ , 0 ) ) = 0 .   □
Remark 20.
In the set-up of Theorem 9, assume α 0 ( mod k + 1 ) and α 0 ( mod k 1 ) .
Set e ˜ 1 : = α k + 1 k a k 1 and f ˜ 1 : = α k + 1 + a k 1 . The proof of Theorem 9 gives the equalities h i ( I 2 A 2 S ( t ̲ , 1 ) ) = 0 , i = 0 , 1 , for a general S S ( Y , e ˜ 1 + f ˜ 1 ) .
Proof of Theorem 6.
It is sufficient to mimic the proof of Theorem 7 using Theorem 9 and lemma similar to lemma 19 instead of Theorem 8.   □

7. Examples

In this section we study a few cases with # A 4 in P n and P 1 × P 1 . For P n , we use the group G : = Aut ( P n ) . For P 1 × P 1 , we use G : = Aut ( P 1 ) × Aut ( P 1 ) . For Y, either P n or P 1 × P 1 , we describe the G-orbits of the set S ( Y , a ) for all a 4 .
At the end of the section, we recall the definition of multiple points and give an easy example to see why our tools are useful.
Remark 21.
Take Y = P n , n 2 . Since P n is 2-transitive, for a 2 , the action of G on S ( Y , a ) is transitive. Hence, 2 A 2 S has the Hilbert function of a general S S ( P n , x + a ) for a general S S ( P n , x ) . For a = 3 , 4 , the set S ( P n , a ) has an open orbit (for n = 4 , we need that n 1 ). If A is in the open orbit, then 2 A 2 S has the Hilbert function of a general S S ( P n , x + a ) for a general S S ( P n , x ) .
(a) Take a = 3 . The action has 2 orbits: the set of collinear points and the set of non-collinear ones. The open orbit is the set of all non-collinear ones.
(b) Take a = 4 . Here, the set of collinear points has infinitely many orbits (they are parametrized by the cross ratio of the 4 points). The set of collinear points is a closed and irreducible subset of P n of dimension 2 n + 4 . There are only a finite number of other orbits, and they are described below in (b1) and (b2).
(b1) Assume n = 2 . The open orbit is formed by sets such that no 3 of the points are collinear. The other orbit is formed by the sets A such that # ( L A ) = 3 for some line L.
(b2) Assume n > 2 . The open orbit is formed by 4 linearly independent points. Another orbit is formed by the sets A S ( P n , 4 ) contained in a plane but such that no 3 of the points are collinear. The other orbit is formed by the sets A such that # ( L A ) = 3 for some line L.
Remark 22.
Take d = 2 , a 2 , x 0 , A S ( P n , a ) and a general S S ( P n , x ) .
Set m : = dim A . The singular locus of a quadric hypersurface is a linear space. Thus h 0 ( I 2 A 2 S ( 2 ) ) = max { 0 , n m x } .
Example 1.
Take d = 3 , a { 3 , 4 } , x 0 , A S ( P n , a ) and a general S S ( P n , x ) with the convention S = if x = 0 .
(a) Assume A L for some line L. We have deg ( 2 A L ) = 2 a , and hence, h 1 ( L , I 2 A L ( 3 ) ) = 2 a 4 . Remark 2 gives h 1 ( I 2 A ( 3 ) ) 2 a 4 . We will check in steps (a1) and (a2) that h 1 ( I 2 A ( 3 ) ) = 2 if a = 3 and h 1 ( I 2 A ( 3 ) ) = n + 3 if a = 4 . We will also see that h 1 1 ( I 2 B ( 3 ) ) = 2 + ( n + 1 ) ( b 3 ) for any line L and any B S ( L , b ) with b 3 .
(a1) Assume n = 2 . Since | I 2 A | is the set of all plane cubics 2 L R with R a line, we have h 0 ( I 2 A 2 S ( 3 ) ) = 3 (resp. 0) if x = 0 (resp. x 1 ). Since h 0 ( I 2 A ( 3 ) ) = 3 , we have h 1 ( I 2 A ( 3 ) ) = 3 a 7 .
(a2) Assume n > 2 and take a general hyperplane H L . The case n = 2 and Remark 2 give h 1 ( I 2 A ( d ) ) 3 a 7 . Consider the residual exact sequence of H:
0 I A ( 2 ) I 2 A ( 3 ) I H 2 A , H ( 3 ) 0
Since h 1 ( I A ( 2 ) ) = 0 if a = 3 and h 1 ( I A ( 2 ) ) = 1 is a = 4 , the long cohomology exact sequence of (16) gives h 1 ( I 2 A ( 3 ) ) = 2 if a = 3 . Note that for a = 3 , L is contained in the base locus of I A ( 2 ) . We find that for a = 3 , all cubic hypersurfaces containing 2 A are singular at all points of L. Hence, for all p L A , we have h 0 ( I 2 A 2 p ( 3 ) ) = h 0 ( I 2 A ( 3 ) ) . Hence, for a = 4 , we have h 1 ( I 2 A ( 3 ) ) = n + 3 .
(b) Assume a = 4 and the existence of a line L such that # ( A L ) = 3 .
(b1) Assume n = 2 . Case (a1) gives h 0 ( I 2 A ( 3 ) ) = 0 . Hence, h 1 ( I 2 A ( 3 ) ) = 2 .
(b2) Assume n > 2 . We want to prove by induction on n that h 1 ( I 2 A ( 3 ) ) = 2 . Let H be a general hyperplane containing the plane A . We have h 1 ( I A ( 3 ) ) = 0 . Hence, it is sufficient to use the long cohomology exact sequence of (16).
(c) Assume n > 2 , the non-existence of a line L such that # ( A L ) = 3 and that M : = A is a plane. We have h 0 ( M , I 2 A M , M ( 3 ) ) = 0 , and hence, h 1 ( M , I 2 A M , M ( 3 ) ) = 2 . Then, as in step (b2), by induction on n, we obtain h 1 ( I 2 A ( 3 ) ) = 2 .
Proposition 4.
Take A S ( P 2 , 4 ) and a general S S ( P 2 , x ) .
(a) Assume that A is contained in a line.
We have h 1 ( I 2 A ( 6 ) ) = 1 , h 1 ( I 2 A ( 5 ) ) = h 1 ( I 2 A 2 S ( 5 ) ) = 2 for x 4 , h 0 ( I 2 A 2 S ( 6 ) ) = 1 and h 1 ( I 2 A 2 S ( 6 ) ) = 2 for x = 5 , h 0 ( I 2 A 2 S ( 6 ) ) = 0 if and only if x 6 . Moreover, we have h 0 ( I 2 A 2 S ( 5 ) ) = 0 if and only if x 4 .
(b) Assume that A is not contained in a line. We have h 0 ( I 2 A 2 S ( 6 ) ) = 0 if x 6 and h 1 ( I 2 A 2 S ( 6 ) ) = 0 if x 5 . We have h 1 ( I 2 A 2 S ( 5 ) ) = h 0 ( I 2 A 2 S ( 5 ) ) = 0 for x = 3 .
Proof. 
Take A contained in a line L. We have deg ( L 2 A ) = 8 , and hence, we obtain h 1 ( L , I 2 A L , L ( 6 ) ) = 1 and h 1 ( L , I 2 A L , L ( 5 ) ) = 2 . Remark 2 gives h 1 ( I 2 A ( 6 ) ) > 0 and h 1 ( I 2 A ( 5 ) ) 2 . Since Res L ( 2 A ) = A and h 1 ( I A ( x ) ) = 0 for all x 3 , the residual exact sequence of L gives h 1 ( I 2 A ( 6 ) ) = 1 , h 1 ( I 2 A ( 5 ) ) = 2 and h i ( I 2 A 2 S ( 6 ) ) = h i ( I A 2 S ( 5 ) ) , i = 0 , 1 . If x 4 , we have h 1 ( I 2 S ( 4 ) ) = 0 by the Alexander–Hirschowitz theorem. Hence, the residual exact sequence of L gives h 1 ( I 2 A 2 S ( 6 ) ) = h 1 ( I 2 A ( 6 ) ) = 2 for x 4 . Take x = 5 and let D be the only conic containing S. Using the residual exact sequence of 2 L , we obtain | I 2 A 2 S ( 6 ) | = { 2 L 2 D } . Hence, h 1 ( I 2 A 2 S ( 6 ) ) = 2 . Adding a general point, we obtain h 0 ( I 2 A 2 S ( 6 ) ) = 0 if x 6 . The residual sequence of 2 L gives h 0 ( I 2 A 2 S ( 5 ) ) = 0 if and only if x 4 .
Now, assume that A is not contained in a line. By the semicontinuity theorem for cohomology (ref. [], Th. III.12.8) and Lemma 4, it is sufficient to use the case in which there is a line L such that # ( L A ) = 3 . Set E : = L A and { o } : = A E . Since h 0 ( O P 2 ( 6 ) ) = 28 , to prove the part on O P 2 ( 6 ) , it is sufficient to prove that h 0 ( I 2 A 2 S ( 6 ) ) = 1 for a general S S ( P 2 , 5 ) . Take p L E and a general S 1 S ( P 2 , 4 ) . We have h i ( L , I 2 A { p } , L ( 6 ) ) = 0 , i = 0 , 1 . By the Differential Horace Lemma applied to p to prove that h 1 ( I 2 A 2 S ( 6 ) ) = 0 , it is sufficient to prove that h 1 ( I A ( 2 p , L ) 2 S 1 2 o ( 5 ) ) = 0 . Take a general q L and a general S 2 S ( P 2 , 3 ) ) . Note that h i ( L , I A ( 2 p , L ) { q } , L ( 5 ) ) = 0 , i = 0 , 1 . Applying the Differential Horace Lemma to q, we find that to prove that h 1 ( I A ( 2 p , L ) 2 S 1 2 o ( 5 ) ) = 0 , it is sufficient to prove that h 0 ( I 2 o 2 S 2 ( 2 q , L ) ( 4 ) ) = 1 . Since S 2 is general and q is general in q, there is a unique conic D containing S 2 { o , q } and it is not tangent to L at q. Thus Res D ( 2 o 2 S 2 ( 2 q , L ) ) = { o , q } S 2 , and hence, we obtain that | I 2 o 2 S 2 ( 2 q , L ) ( 4 ) | = { 2 D } .
Let U denote the set of all A S ( P 2 , 4 ) such that there is a line L with # ( A L ) = 3 . The set U is a G-orbit. Hence, h i ( I 2 S 2 A ( 5 ) ) , i = 0 , 1 , may be computed in the following way. Set { o } : = A L A . By the residual exact sequence of L for i = 0 , 1 , we have the equality h i ( I 2 A 2 ( 5 ) ) = h i ( I 2 o 2 S ( S L ) ( 4 ) ) with { o } S , a general element of S ( P 2 , 4 ) ) (and hence h 1 ( I 2 A 2 o ( 4 ) ) = 0 and h 0 ( I 2 o 2 S ( 4 ) ) = 3 by the Alexander–Hirschowitz theorem) and 3 general points of a general line of P 2 . Thus, h i ( I 2 o 2 S ( S L ) ( 4 ) ) = 0 , i = 0 , 1 .   □
Remark 23.
Take Y = P 1 × P 1 . For any A S ( Y , a ) , a 2 , we have the pair ( # π 1 ( A ) , # π 2 ( A ) ) of positive integers and ( # π 1 ( A ) , # π 2 ( A ) ) = ( # π 1 ( B ) , # π 2 ( B ) ) if A and B are contained in the same G-orbit. Not all pairs of integers ( e , f ) with 1 e a and 1 f a occur as ( # π 1 ( A ) , # π 2 ( A ) ) for some S ( Y , a ) . For instance, if # π i ( A ) = 1 , then # π 3 i ( A ) = a .
Note that h 0 ( I 2 A ( 1 , 1 ) ) = 0 for all a 2 and all A S ( Y , a ) .
(a) The variety S ( Y , 2 ) has 3 G-orbits, the open one, the one with points with the same first projection and the one with the same second projection.
(b) The variety S ( Y , 3 ) has 5 G-orbits, distinguished by the pair ( # π 1 ( A ) , # π 2 ( A ) ) .
(c) The set S ( Y , 4 ) has infinitely many orbits, even among the ones contained in a curve of multidegree ϵ 1 or a curve of multidegree ϵ 2 . Since Aut ( P 1 ) is 3-transitive, there is a unique G-orbit associated to ( 3 , 2 ) and a unique G-orbit associated to ( 2 , 3 ) . There are 2 different G-orbits associated to ( 2 , 2 ) . One of them is formed by the elements of S ( Y , 4 ) , which are the complete intersection of a reduced curve of multidegree ( 2 , 0 ) and a reduced curve of multidegree ( 0 , 2 ) . The other one is formed by the set of all A S ( Y , 4 ) contained in the smooth locus of a reducible element of | O Y ( 1 , 1 ) | . The classical notion of cross-ratio shows the existence of a 2-dimensional family of orbits with pair ( 4 , 4 ) . We have h 0 ( I A ( 1 , 1 ) ) = 2 if and only if there is i { 1 , 2 } and A is contained in a curve of multidegree ϵ i . Note that 4 different orbits are formed by sets A with h 0 ( I A ( 1 , 1 ) ) = 1 .
Notation 1.
For any smooth manifold M, any p M and any positive integer m, let ( m p , M ) denote the closed subscheme of M with ( I p ) m as its ideal sheaf. If n : = dim M , then deg ( ( m p , M ) ) = n + m 1 n . Set ( 0 p , M ) : = . For all finite subsets of M, set ( m S , M ) = p S ( m p , M ) . Write m p and m S instead of ( m p , M ) and ( m S , M ) , respectively, if M = Y .
The zero-dimensional scheme ( m p , M ) is called a multiple point of M of multiplicity m. L. Chiantini and Th. Mawking considered the postulation of the union of a triple point and finitely many general double points [,]. The next result considers the same problem for 2 multiple points of P 1 × P 1 , the ones corresponding to one of the smaller orbits in part (a) of Remark 23.
Proposition 5.
Take Y = P 1 × P 1 and A S ( Y , 2 ) such that # π 1 ( A ) ) = 1 . Let L be the only element of | O Y ( 0 , 1 | containing A. For all positive integers m, d and all zero-dimensional schemes Z such that Z L = , we have h 0 ( I m A Z ( 1 , d ) ) = h 0 ( I Z ( 1 , d m ) ) .
Proof. 
Note that deg ( x A L ) = 2 x for all integers x 0 and that deg ( O L ( 1 , t ) ) = 1 for all t Z . Thus, it is sufficient to use m times the residual exact sequence of L.   □
For another case of multiple points with a fixed support, see [].

8. Discussion

Motivated by its relation with the computation of the dimensions of the secant varieties of an embedded variety X, we consider an interpolation problem for unions of a + x double points, a of them prescribed, while the other x are general in X. Our results are in cases of Veronese embeddings of projective spaces and of Segre–Veronese embeddings of multiprojective spaces. The multiprojective spaces we consider are the ones with all factors of dimension 1 and the ones with 2 factors, one of them one-dimensional. We use the known fundamental results for general unions of double points in these cases [,,,]. We suggest that other places to be explored are Grassmannians and the Schur factors of tensors [,,]. We proposed a related problem when a of the double points are allowed to vary in a prescribed subvariety. This related problem often comes in some inductive proofs on the interpolation of general double points (the so-called Horace Method). We suggest to explore here even when instead of a subvariety, we have a flag of subvarieties and prescribe the number of double points whose reduction is contained in each stratum. In the last section, we gave several examples on how to use the tools of the paper for related problems, even with points of higher multiplicities.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Ådlandsvik, B. Joins and higher secant varieties. Math. Scand. 1987, 61, 213–222. [Google Scholar] [CrossRef]
  2. Bernardi, A.; Carlini, E.; Catalisano, M.V.; Gimigliano, A.; Oneto, A. The Hitchhiker guide to: Secant varieties and tensor decomposition. Mathematics 2018, 6, 314. [Google Scholar] [CrossRef]
  3. Landsberg, J.M. Tensors: Geometry and Applications; Graduate Studies in Mathematics; American Mathematical Society: Providence, RI, USA, 2012; Volume 128. [Google Scholar]
  4. Abo, H.; Brambilla, M.C. On the dimensions of secant varieties of Segre-Veronese varieties. Ann. Mat. Pura Appl. 2013, 192, 61–92. [Google Scholar] [CrossRef]
  5. Abo, H.; Brambilla, M.C.; Galuppi, F.; Oneto, A. Non-defectivity of Segre-Veronese varieties. Proc. Am. Math. Soc. Ser. B 2024, 11, 589–602. [Google Scholar] [CrossRef]
  6. Galuppi, F.; Oneto, A. Secant non-defectivity via collisions of fat points. Adv. Math. 2022, 409, 108657. [Google Scholar] [CrossRef]
  7. Laface, A.; Postinghel, E. Secant varieties of Segre-Veronese embeddings of ( P 1 )r. Math. Ann. 2013, 356, 1455–1470. [Google Scholar] [CrossRef]
  8. Alexander, J.; Hirschowitz, A. Un lemme d’Horace différentiel: Application aux singularité hyperquartiques de P5. J. Algebr. Geom. 1992, 1, 411–426. [Google Scholar]
  9. Alexander, J.; Hirschowitz, A. La méthode d’Horace éclaté: Application à l’interpolation en degré quatre. Invent. Math. 1992, 107, 585–602. [Google Scholar] [CrossRef]
  10. Alexander, J.; Hirschowitz, A. Polynomial interpolation in several variables. J. Algebr. Geom. 1995, 4, 201–222. [Google Scholar]
  11. Brambilla, M.C.; Ottaviani, G. On the Alexander-Hirschowitz Theorem. J. Pure Appl. Algebra 2008, 212, 1229–1251. [Google Scholar] [CrossRef]
  12. Chandler, K. Uniqueness of exceptional singular quartics. Proc. Am. Math. Soc. 2004, 132, 347–352. [Google Scholar] [CrossRef]
  13. Chandler, K. A brief proof of a maximal rank theorem for generic double points in projective space. Trans. Am. Math. Soc. 2000, 353, 1907–1920. [Google Scholar] [CrossRef]
  14. Chandler, K. Linear systems of cubics singular at general points of projective space. Compos. Math. 2002, 134, 269–282. [Google Scholar] [CrossRef]
  15. Postinghel, E. A new proof of the Alexander-Hirschowitz interpolation theorem. Ann. Mat. Pura Appl. 2012, 191, 77–94. [Google Scholar] [CrossRef]
  16. Hartshorne, R. Algebraic Geometry; Springer: Berlin/Heidelberg, Germany, 1977. [Google Scholar]
  17. Alexander, J.; Hirschowitz, A. An asymptotic vanishing theorem for generic unions of multiple points. Invent. Math. 2000, 140, 303–325. [Google Scholar] [CrossRef]
  18. Ballico, E.; Bernardi, A.; Mańdzuk, T. Tensoring by a plane preserves secant-regularity in degree at least two. Ann. Mat. Pura Appl. 2025, 204, 489–511. [Google Scholar] [CrossRef]
  19. Ballico, E. On the non-defectivity of Segre-Veronese embeddings. Math. Z. 2024, 308, 6. [Google Scholar] [CrossRef]
  20. Abo, H.; Brambilla, M.C. Secant varieties of Segre-Veronese varieties P m × P n embedded by O (1,2). Exp. Math. 2009, 18, 369–384. [Google Scholar] [CrossRef]
  21. Abo, H. On non-defectivity of certain Segre-Veronese varieties. J. Symb. Comput. 2010, 45, 1254–1269. [Google Scholar] [CrossRef]
  22. Abo, H.; Brambilla, M.C. New examples of defective secant varieties of Segre-Veronese varieties. Collect. Math. 2012, 63, 287–297. [Google Scholar] [CrossRef]
  23. Chiantini, L.; Markwig, T. Triple-point defective ruled surfaces. J. Pure Appl. Algebra 2008, 212, 1337–1346. [Google Scholar] [CrossRef]
  24. Chiantini, L.; Markwig, T. Triple-point defective surfaces. Adv. Geom. 2010, 10, 527–547. [Google Scholar] [CrossRef]
  25. Bocci, C.; Chiantini, L. The effect of points fattening on postulation. J. Pure Appl. Algebra 2011, 21, 89–98. [Google Scholar] [CrossRef]
  26. Galgano, V. Identifiability and singular locus of secant varieties to spinor varieties. Commun. Contemp. Math. 2025, 2550089. [Google Scholar] [CrossRef]
  27. Galgano, V.; Staffolani, R. Identifiability and singular locus of secant varieties to Grassmannians. Collect. Math. 2025, 76, 105304. [Google Scholar] [CrossRef]
  28. Staffolani, R. Schur apolarity. J. Symb. Comput. 2023, 114, 37–73. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.