Home The geometry of rank drop in a class of face-splitting matrix products: Part I
Article Open Access

The geometry of rank drop in a class of face-splitting matrix products: Part I

  • Erin Connelly , Sameer Agarwal , Alperen Ergur and Rekha R. Thomas EMAIL logo
Published/Copyright: August 5, 2024
Become an author with De Gruyter Brill

Abstract

Given k points (xi, yi) ∈ ℙ2 × ℙ2, we characterize rank deficiency of the k × 9 matrix Zk with rows xiyi in terms of the geometry of the point configurations {xi} and {yi}. In this paper we present results for k ≤ 6. For k ≤ 5, the geometry of the rank-drop locus is characterized by cross-ratios and basic (projective) geometry of point configurations. For the case k = 6 the rank-drop locus is captured by the classical theory of cubic surfaces. The results for k = 7, 8 and 9 are presented in the sequel [7] to this paper.

MSC 2010: 14Q10; 68T45

1 Introduction

Let ⊗ denote the Kronecker product; see [30]. We are interested in solving the following problem:

Problem 1.1

Given k points (xi, yi) ∈ ℝ3 × ℝ3 where k ≤ 9, consider the k × 9 matrix Zk whose rows are xiyi for i = 1, …, k, i.e.,

Zk=x1y1xkyk.

Delineate the geometry of point configurations {xi} and {yi} for which rank(Zk) < k.

In signal processing, the matrix Zk is known as the face-splitting product [27] of the two matrices Xk ∈ ℝk×3 with rows x1,,xk and Yk ∈ ℝk×3 with rows y1,,yk. The face-splitting product is a special case of the Khatri–Rao matrix product in linear algebra and statistics [22]. However, our interest in Problem 1.1 comes from an estimation problem in 3D computer vision known as the 7-point problem [18].

Problem 1.2

(7-point problem). Given 7 points (i, ŷi) ∈ ℝ2 × ℝ2, find a matrix F ∈ ℝ3×3 such that det(F) = 0 and

yiFxi=0fori=1,,7,wherexi=x^i1andyi=y^i1. (1)

In 3D computer vision the world (or the scene) is modeled as a collection of points in ℙ3. Pinhole cameras imaging the scene are modeled as full rank linear projections from ℙ3 to ℙ2.

Let P1, …, P7 ∈ ℙ3 be a scene consisting of seven world points, and let A1 and A2 be two pinhole cameras such that they map the scene to images points x1, …, x7 and y1, …, y7 respectively (here we assume that the images are finite points in ℙ2, so they can be de-homogenized to points in ℝ2). A solution F to the 7-point problem that has rank 2 is known as a fundamental matrix. The fundamental matrix determines the pair of cameras A1 and A2 up to a homography of ℙ3; see [18; 23].

Observe that the linear constraints in Problem 1.2can be re-written as [1]

(xiyi)vec(F)=0,for i=1,,k,

where vec(F) is the 9-dimensional vector obtained by concatenating the columns of F. Thus solving the 7-point problem requires computing the intersection of the null space of a 7 × 9 face-splitting product with a determinantal variety. Generically it will have three distinct solutions, one of which will always be real. When this is not the case, i.e., when Problem 1.2has less than three distinct solutions or has an infinite number of solutions, it is said to be ill-posed [10].

Besides being interesting on its own as a mathematical phenomenon, ill-posedness is of practical interest because of a meta-theorem in numerical analysis that says that the numerical conditioning of a problem is inversely related to the distance to the nearest ill-posed problem [9; 8; 5]. So, understanding the data for which Problem 1.2is ill-posed also helps us understand how and when it is numerically hard to solve it using finite precision arithmetic.

There are two ways in which Problem 1.2can have an infinite number of solutions. The first is that the null space of Z7 has dimension greater than 2. The second is that the null space even though it is 2-dimensional, lies entirely inside the determinantal variety. So, a sufficient condition for ill-posedness is that the corresponding face-splitting product matrix be rank deficient and thus our interest in Problem 1.1.

A first answer to Problem 1.1 is that rank(Zk) < k if and only if all the maximal minors of Zk are zero. These minors are bi-homogeneous polynomials in xi and yi, and typically do not shed much light on the geometry of the point configurations {xi} and {yi} that cause Zk to drop rank. Our goal is to characterize the rank deficiency of Zk in terms of the geometry of {xi} and {yi}.

To this end, we begin by observing that the rank of Zk is a projective invariant. First, note that multiplying the xi’s and yi’s by non-zero scalars scales the rows of Zk and does not change its rank. Second, if H1, H2 are invertible matrices, then from the mixed-product property of Kronecker products we have that

Zk=(H1x1)(H2y1)(H1xk)(H2yk)=(x1y1)(H1H2)(xkyk)(H1H2)=Zk(H1H2). (2)

The matrix H1H2 has full rank since H1 and H2 are invertible, and hence rank(Zk)=rank(Zk). These observations imply that rank(Zk) is independent of scaling and choice of coordinates in ℝ3 × ℝ3. Therefore, we should be studying the problem over projective space. While our computer vision applications care about real input, mathematically there is no loss in assuming that the input points (xi, yi) are complex because ℝ ⊂ ℂ. Therefore, letting ℙ denote projective space over ℂ we pass to the following formulation of Problem 1.1 which will be the main subject of study in this paper.

Problem 1.3

Given k points (xi, yi) ∈ ℙ2 × ℙ2 where k ≤ 9, consider the k × 9 matrix Zk whose rows are xiyi for i = 1, …, k, i.e.,

Zk=x1y1xkyk.

Delineate the geometry of point configurations {xi} and {yi} for which rank(Zk) < k.

Let Zk{l} denote a submatrix of Zk with l of the k rows. If rank(Zk{l}) < l, then rank(Zk) < k, and any condition on the l points (xi, yi) that contribute the rows of Zk{l} and cause Zk{l} to drop rank will be a condition on {(xi, yi), i = 1, …, k} such that Zk drops rank. We will call such a condition an inherited condition for rank drop. For each value of k we will be most interested in the non-inherited conditions on the input that make Zk rank deficient.

As an illustration of the type of answers we will provide, consider the version of Problem 1.3 in ℙ1 × ℙ1. Suppose {xi} and {yi} are each sets of distinct points in ℙ1. In Theorem 4.1 we prove that the 4 × 4 matrix Z4 is rank deficient if and only if the cross-ratio of x1, x2, x3, x4 equals that of y1, y2, y3, y4. The cross-ratio is the fundamental projective invariant of four points in ℙ1.

It is a classical fact that the geometry of a finite set of points in ℙ2, that remains invariant under the action of PGL(3), can be expressed via polynomials in the 3 × 3 determinants (brackets) of the matrices whose rows are the points. Many of our answers to Problem 1.3 will be phrased in terms of these bracket expressions. The results have analogs in terms of cross-ratios when the points are in general position.

We solve Problem 1.3 across this paper (Part I) and its sequel (Part II, see [7]). In this paper we address the Cases 2 ≤ k ≤ 6 and in the next we address k = 7, 8, 9. A great deal of classical algebraic geometry emerges starting with k = 6. This motivated the split since k = 6 is rather elaborate and we wanted to keep the papers to a reasonable length. Also, the flavors of results in the two papers are slightly different. For k ≤ 6, we are able to obtain both geometric and algebraic characterizations of rank drop without any assumptions on the points. For k ≥ 7, the tools and results become more geometric and require a type of genericity of the input.

Related work

The study of ill-posedness of Problem 1.2 has a long history in computer vision.

One prominent direction is the work on critical configurations. A configuration (or a reconstruction) is a set of cameras and world points in ℙ3. It is said to be critical if there exist other projectively non-equivalent configurations that produce the same images [18; 21; 2; 3]. There is no simple way to go from these configurations to their images, which is what we need to study the ill-posedness of Problem 1.2.

Perhaps the most well known result on the ill-posedness of Problem 1.2 is that if {xi} and {yi} are related by a homography of ℙ2 then Problem 1.2 has an infinite number of solutions [23]. This is far from a complete result as it is only a sufficient condition.

Fan et al. study conditioning of Problem 1.2 using a definition based on the world scene [15, Definition 1]. Using this definition allows them to write down a simple formula for the condition number in [15, Proposition 2]. Their world-space result [15, Theorem 2] on the ill-posed locus of Problem 1.2 has substantial overlap with the results on critical configurations [17; 18; 3]. Their result in image-space [15, Theorem 4], which is what is relevant to this paper, is algebraic. It is stated in terms of the existence of a certain high degree polynomial that vanishes on {xi} and {yi} as opposed to describing their geometry.

Bertolini, Magri and Turrini have studied critical configurations in multiple settings. The paper closest to our work is [1] which studies critical configurations for 2-views in terms of {xi} and {yi}. They show that if the images admit multiple real reconstructions then there is a quadratic Cremona transformation taking xiyi. This is also the content of Lemma 3.6 in our paper [7], where we develop this connection further.

Summary of results and organization of the paper

In Section 2 we show that one can make some simplifying assumptions on the input to Problem 1.3 without any loss of generality. While these assumptions are not needed for the theory, they are helpful and sometimes critical for computations. In Section 3 we review the tools from invariant theory that are needed in this paper.

In Section 4 we present our results for k ≤ 4. Theorem 4.2 solves Problem 1.3 for k = 2, 3, 4. Section 5 considers k = 5 and presents two theorems. Theorem 5.1 characterizes the new conditions for rank drop in Z5 in terms of bracket equations (19). A consequence is that when k = 5, a necessary condition for rank drop is that one set of points (say the yi’s) is on a line . The second result is Theorem 5.2 which gives a geometric characterization of rank deficiency: when the yi’s on the line are distinct and the xi’s are in general position, then the rank drop of Z5 is equivalent to the existence of an isomorphism from the conic ω through the xi’s to the line , taking xi to yi.

Finally, we consider the case k = 6 in Section 6 where we will see that Problem 1.3 is intimately related with the theory of cubic surfaces in ℙ3. An important new fact here is that given 5 points (xi, yi) in general position there is a unique 6th point (x6, y6) that makes Z6 rank deficient. This 6th point admits a rational expression in terms of the 5 points (Lemma 6.1).

The results in this section are organized in two parts. In Subsection 6.1 we characterize the rank deficiency of Z6 using brackets; Theorem 6.5 covers the case of neither the xi’s nor the yi’s being in a line, and Theorem 6.8 addresses the situation in which one set of points is in a line. When the points are in general position, we identify a simple algebraic check (Theorem 6.7) for rank deficiency in terms of a vector of classical invariants (39) of 6 points in ℙ2.

In Subsection 6.2, we explain the origins of the algebraic theorems in Subsection 6.1 using geometry. The central result here Theorem 6.10 which says that when the points are in general position, then rank(Z6) < 6 if and only if there is a smooth cubic surface 𝓢 that arises as the blow up of ℙ2 in both the x points and the y points such that the two sets of exceptional curves form a Schläfli double six on 𝓢. This result allows for a simple determinantal representation of 𝓢 using the null space of Z6.

We then interpret Theorem 6.10 in two different ways. The first is Theorem 6.21 that identifies the unique 6th point, given 5 points in general position, that forces Z6 to drop rank. The construction in this theorem relies on a recipe due to Sturm via conics in ℙ2 × ℙ2, and this result proves the algebraic Theorem 6.5.

Theorem 6.10 can be rephrased in yet another way using the Cremona hexahedral form of a cubic surface (Theorem 6.22). This result explains the algebraic Theorem 6.7. Using the invariants needed in this theorem along with the classical Joubert invariants of 6 points in a line, we complete the answer to Problem 1.3 when k = 6 by proving Theorem 6.8.

A note about computations

All of the algebraic statements in this paper are proven using the computational algebra software package Macaulay2; see [16]. The codes that we used can be found at github.com/rekharthomas/rankdrop.

A basic strategy we employ to understand rank drop is to use Macaulay2 to decompose the ideal of maximal minors of Zk over ℚ into its associated primes (over ℚ). Then we interpret these ideals geometrically to state the conditions for rank drop. These geometric conditions will hold over ℝ and ℂ as well since the prime decomposition over ℚ can only refine over the larger fields.

2 Preprocessing the input

We begin by showing that one can make simplifying assumptions on the input to Problem 1.3 which is a collection of k ≤ 6 points (xi, yi) ∈ ℙ2 × ℙ2.

We sometimes refer to (xi, yi) as a point pair, especially when we need to decouple the components and think of the configurations xi and yi separately. It will also be convenient to refer to the ℙ2 containing {xi} as Px2 and the ℙ2 containing {yi} as Py2 . Let PGL(3) be the group of invertible 3 × 3 matrices over ℂ up to scaling. An element H ∈ PGL(3) induces a homography (projective isomorphism) on ℙ2.

In the introduction we informally showed that the rank of Zk = (xiyi)i=1k is invariant under PGL(3) × PGL(3) acting on Px2 × Py2 by left multiplication. This allows us to change bases in Px2 and Py2 independently. We state the result formally below:

Lemma 2.1

Let (xi, yi) ∈ ℙ2 × ℙ2 for i = 1, …, k and let H1, H2 ∈ PGL(3) be homographies acting on Px2 and Py2 respectively, i.e. (x, y) ↦ (H1x, H2y). Then Zk = (xiyi)i=1k has the same rank as Zk=((H1xi)(H2yi))i=1k.

Lemma 2.1 allows us to make certain assumptions on the configurations {xi} and {yi} which we phrase as corollaries to the lemma. While our proofs do not need these assumptions, they become critical for computations in Macaulay2. The first assumption is that the xi’s and yi’s in the input to Problem 1.3 are finite points. With a view towards computations, in the following lemma and in several parts of the paper, we fix representatives of points in ℙ2 and write them out as vectors in ℂ3.

Corollary 2.2

Without loss of generality, we can assume that the input to Problem 1.3 is of the form xi = (xi1, xi2, 1) and yi = (yi1, yi2, 1) for all i = 1, …, k, where xi1, xi2, yi1, yi2 ∈ ℂ. Hence

xiyi=(xi1yi1,xi1yi2,xi1,xi2yi1,xi2yi2,xi2,yi1,yi2,1). (3)

Proof

Let xi Px2 and yi Py2 be k pairs of points where k is finite, and suppose that some of these points lie on the line at infinity in their ℙ2. Since we have only finitely many points, there exists a line , common to both ℙ2, which contains none of these points. Let H be a homography that sends to the line at infinity in both ℙ2. Then Hxi and Hyi are finite points for all i. By Lemma 2.1, the rank of Zk = (xiyi)i=1k equals the rank of Zk=((Hxi)(Hyi))i=1k and hence we can work with the transformed points to understand the rank deficiency of Zk.□

The other assumption we can make is that sufficiently generic configurations allow some of the points to be fixed, which will prove useful for computations.

Corollary 2.3

When no three of the xis or yis are collinear in their respective2, we may fix the first four in each set to be

x1=y1=(0,0,1),x2=y2=(1,0,1),x3=y3=(0,1,1),x4=y4=(1,1,1). (4)

At certain times, for the sake of symmetry, we will instead fix the first four points as

x1=y1=(1,0,0),x2=y2=(0,1,0),x3=y3=(0,0,1),x4=y4=(1,1,1). (5)

Since the two fixings are related by a homography, results from one fixing transfer to the other.

3 Cross-ratios and brackets

By Lemma 2.1, the rank deficiency of Zk is invariant under the action of PGL(3) × PGL(3) on ℙ2 × ℙ2 given by (H1, H2) ⋅ (x, y) ↦ (H1x, H2y). Therefore, the input to Problem 1.3 can be thought of as the PGL(3)-orbits of k points in each ℙ2. The intrinsic geometry of k points p1, …, pk ∈ ℙ2 are the properties of the configuration that remain intact/invariant under the action of PGL(3). Since an element of PGL(3) can be assumed to have determinant 1, the above geometry is precisely the invariant properties, under the action of SL(3), of the configuration obtained by replacing each pi ∈ ℙ2 with a representative in ℂ3. We now introduce the basic tools from the invariant theory of point sets needed to formulate our results. There are numerous references for this material such as [6], [11], [12], [13] and [29].

Let ℂ[p] := ℂ[pi1, pi2, pi3, i = 1, …, k] be the polynomial ring in the variables pij for i = 1, …, k, j = 1, 2, 3, that represent the coordinates of k points p1, …, pk in ℙ2 written via representatives in ℂ3. A polynomial f ∈ ℂ[p] is invariant under SL(3) if f(p1, …, pk) = f(Hp1, …, Hpk) for all H ∈ SL(3). It is a well-known classical result that the invariant ring ℂ[p]SL(3), which is the collection of all polynomials in ℂ[p] that are invariant under SL(3), is generated by the 3 × 3 minors of the matrix whose rows are the symbolic p1,,pk. In other words, ℂ[p]SL(3) is generated as an algebra by the bracket polynomials

[ijl]:=detpi1pi2pi3pj1pj2pj3pl1pl2pl3 (6)

for all distinct choices of i, j, l ∈ {1, …, k}. In our situation, we will write [ijl]x (respectively, [ijl]y) when the input points come from {xi} (respectively, {yi}). Equations in brackets can be used to express geometry. For example, three points pi, pj, pl ∈ ℙ2 are collinear if and only if [ijl] = 0, and 6 points lie on a conic if and only if they satisfy the bracket equation (12) in Lemma 3.5.

Often we consider points in a line in ℙ2 as points in ℙ1 where we can compute the bracket [ij] for pi, pj by identifying with the line parameterized as say (*, *, 0), and then dropping the 0 coordinate to get

[ij]:=detpi1pi2pj1pj2. (7)

When we have many points p1, …, pk on a line and a point u not on that line, the ℙ2 brackets [iju] will be closely related to the ℙ1 brackets [ij].

Lemma 3.1

For points p1, …, pk on a line and a point u not on the line, there exists a nonzero scalar λ depending only on the choice of coordinate representative of u such that [iju] = λ[ij] for all i, j.

Proof

We can assume without loss of generality that pi = (pi1, pi2, 0) for all i. We then calculate

[iju]=detpi1pi20pj1pj20u1u2u3=u3[ij].

Definition 3.2

The cross-ratio of an ordered list of 4 points p1, …, p4 ∈ ℙ1 is

(p1,p2;p3,p4):=[13][24][14][23]. (8)

The cross-ratio is the only projective invariant of 4 points in ℙ1 in the following sense.

Lemma 3.3

If p1, …, p4 ∈ ℙ1 and q1, …, q4 ∈ ℙ1 are two sets of points, then (p1, p2; p3, p4) = (q1, q2; q3, q4) if and only if there exists a homography H : ℙ1 → ℙ1 such that Hpi = qi for all i.

Permuting the points on the line changes the cross-ratio in a systematic way, i.e., if

(p1,p2;p3,p4)=(q1,q2;q3,q4)

then

(pσ(1),pσ(2);pσ(3),pσ(4))=(qσ(1),qσ(2);qσ(3),qσ(4))

for all σS4. Thus we do not need to consider multiple orderings. The cross-ratio is 0, 1, ∞, or 00 if and only if the points are not distinct. See [26, III, § 4–5].

Example 1

We now give an example to illustrate that equations in brackets can be more robust than equalities of cross-ratios. Consider the following points in ℙ1 where the first set has repetition:

x1=(1,1)x2=(1,1)y1=(8,1)y2=(4,1)x3=(1,1)x4=(3,1)y3=(2,1)y4=(5,1).

Then

(x1,x2;x3,x4)=00,(y1,y2;y3,y4)=1

and therefore the cross-ratios are not equal. However, if we write the cross-ratio equality

(x1,x2;x3,x4)=(y1,y2;y3,y4) (9)

in bracket form as

[13]x[24]x[14]y[23]y=[13]y[24]y[14]x[23]x (10)

then equality does hold.

In Example 1, the points (xi, yi) lie on the variety defined by the bracket equation (10) but do not satisfy the cross ratio equality (9). The discrepancy is because cross-ratios are well-defined only on an open set, while the corresponding bracket equation holds on the Zariski closure of this open set. Since our interest is in describing the algebraic variety of input points that make Z4 rank deficient, the bracket expressions are the correct choice. However, whenever the points are in sufficiently general position, we can pass to cross-ratio expressions which, besides being elegant, have the advantage that they can be visualized easily.

We will also need planar and conic cross-ratios which we now define.

Definition 3.4

The 5 point cross-ratio of p1, …, p5 ∈ ℙ2 is

(p1,p2;p3,p4;p5):=[135][245][145][235]. (11)

This is also called a planar cross-ratio or the cross-ratio around p5.

Note that planar cross-ratios are preserved under a homography of ℙ2. For 5 distinct points, the cross-ratio around x5 can be obtained geometrically by drawing the 4 lines pip5 for i = 1, 2, 3, 4, cutting them with a transversal, and then computing the cross-ratio of the intersection points. See Figure 1. Also, the planar cross-ratio is transformed by permutations of p1, …, p4 similarly to the line cross-ratio.

Figure 1 
For distinct points p1, …, p5 ∈ ℙ2, the planar cross-ratio (p1, p2; p3, p4; p5) equals the 4 point cross-ratio (a1, a2; a3, a4).
Figure 1

For distinct points p1, …, p5 ∈ ℙ2, the planar cross-ratio (p1, p2; p3, p4; p5) equals the 4 point cross-ratio (a1, a2; a3, a4).

Lemma 3.5

([29, Example 3.4.3]). A collection of 6 points p1, …, p6 ∈ ℙ2 lie on a conic if and only if

[135][245][146][236]=[136][246][145][235], (12)

or generically, if the following cross-ratio equality holds:

(p1,p2;p3,p4;p5)=(p1,p2;p3,p4;p6). (13)

Lemma 3.5 makes the following definition of a conic cross-ratio well-defined.

Definition 3.6

The conic cross-ratio of 4 points p1, …, p4 on a (non-degenerate) conic ω is

(p1,p2;p3,p4)ω:=(p1,p2;p3,p4;p) (14)

where pω is any new point.

Any non-degenerate conic ω is isomorphic to ℙ1. If ι : ω → ℙ1 is one such isomorphism then

(p1,p2;p3,p4)ω=(ι(p1),ι(p2);ι(p3),ι(p4)). (15)

4 The cases k ≤ 4

We will now answer Problem 1.3 in the cases k ≤ 4. As a warm up to Problem 1.3 we first consider the analogous question in ℙ1 × ℙ1. Recall that any condition on a proper subset of {(xi, yi)} that makes the corresponding rows of Zk dependent is called an inherited condition for the rank drop of Zk.

Theorem 4.1

Consider the k × 4 matrix Zk = (xiyi)i=1k where (xi, yi) ∈ ℙ1 × ℙ1 and k ≤ 4. Then

  1. rank(Z2) < 2 if and only if x1 = x2 and y1 = y2,

  2. rank(Z3) < 3 if and only if either an inherited condition holds or x1 = x2 = x3 or y1 = y2 = y3,

  3. rank(Z4) < 4 if and only if the following bracket equation holds:

    [13]x[24]x[14]y[23]y=[14]x[23]x[13]y[24]y. (16)

    Generically, x1, …, x4 and y1, …, y4 are distinct and then (16) holds if and only if there is a homography H ∈ PGL(2) such that Hxi = yi for i = 1, …, 4.

Proof

In each of the above cases, Zk will be rank deficient if and only if all its maximal minors vanish. Hence we compute the ideal of k × k minors of the symbolic Zk in each case and decompose it to understand rank deficiency. While computations in Macaulay2 are done over ℚ, our results hold over ℂ as explained at the end of the Introduction. By Corollary 2.2 we can fix the last coordinates of xi and yi to 1.

If k = 2 the ideal of 2 × 2 minors of Z2 is the prime ideal 〈x11x21, y11y21〉 which says that rank drops if and only if x1 = x2 and y1 = y2. We will call this the repeated point condition.

If k = 3, the ideal of 3 × 3 minors of the symbolic Z3 matrix is radical and decomposes into 5 prime ideals. Three of them correspond to repeated points xi = xj and yi = yj for ij. These are inherited for rank drop from the k = 2 case. The remaining 2 prime ideals say that x1 = x2 = x3 and y1 = y2 = y3 respectively.

When k = 4, we can use Macaulay2 to check that

det(Z4)=[14]x[23]x[13]y[24]y[13]x[24]x[14]y[23]y

which yields (16). In this case, the ideal of maximal minors is principal and prime, generated by det(Z4), and does not decompose. However, note that the bracket expression for det(Z4) vanishes under all of the conditions inherited from k = 2 and k = 3. Furthermore, when the xi’s and yi’s are distinct then (16) is equivalent to (x1, x2; x3, x4) = (y1, y2; y3, y4) and the homography follows by Lemma 3.3.□

We now consider Problem 1.3 for k ≤ 4 points (xi, yi) ∈ ℙ2 × ℙ2. Again all of the statements can be checked in Macaulay2 by following the same general strategy as in the proof of Theorem 4.1.

Theorem 4.2

Suppose that (xi, yi) ∈ ℙ2 × ℙ2 and Zk = (xiyi)i=1k for k ≤ 4. Then

  1. rank(Z2) < 2 if and only if x1 = x2 and y1 = y2.

  2. rank(Z3) < 3 if and only if either an inherited condition holds or x1 = x2 = x3 and y1, y2, y3 are on a line, or x1, x2, x3 are on a line and y1 = y2 = y3.

  3. rank(Z4) < 4 if and only if an inherited condition holds or one of the following is true:

    1. x1 = x2 = x3 = x4 or y1 = y2 = y3 = y4.

    2. The points x1, …, x4 Px2 are on a line and y1, …, y4 Py2 are on a line, and

      [13]x[24]x[14]y[23]y=[14]x[23]x[13]y[24]y. (17)

      Generically, the collinear points x1, …, x4 and y1, …, y4 are distinct and then (17) holds if and only if there is a homography H ∈ PGL(3) such that Hxi = yi for i = 1, …, 4.

Proof

The proof of (1) and (2) can be done using Macaulay2 by computing the ideal of maximal minors of Zk and then decomposing the radical of the ideal to get all possible conditions for rank drop. The matrix Zk is written symbolically as in (3) where by Corollary 2.2 we assume that the last coordinates of all xi’s and yi’s are 1. For instance, the ideal of 3 × 3 minors of Z3 is radical and of dimension 8 and degree 3. It is the intersection of 5 prime ideals of which 3 correspond to the repeated point conditions inherited from k = 2. The fourth ideal provides the new condition that x1 = x2 = x3 and y1, y2, y3 are on a line. The fifth ideal provides the analogous condition with the roles of x and y switched. These ideals appear in the form

x11x21,x12x22,x11x31,x12x32,[123]y and y11y21,y12y22,y11y31,y12y32,[123]x.

The ideal of maximal minors of Z4 is radical of dimension 12 and degree 6. It decomposes into 17 primes ideals:

  • 6 of the primes correspond to repeated points (inherited from k = 2). All of these have dimension 12 and degree 1.

  • 8 of the primes correspond to 3 points on one side coinciding and the corresponding points on the other side on a line (inherited from k = 3). All of these ideals have dimension 11 and degree 2.

  • 2 of the primes say x1 = x2 = x3 = x4 and y1 = y2 = y3 = y4 respectively (new conditions for k = 4). These ideals have dimension 10 and degree 1.

  • The last prime component, call it J, has dimension 11 and degree 24. It implies that the xi and yi are on a line (new condition for k = 4). One way to verify this claim is to check that the ideal generated by the brackets [123]x, [124]x and [123]y, [124]y is contained in J. These brackets being 0 encode the collinearity of xi’s and the collinearity of yi’s.

One can understand J more precisely by making the following calculation. If the x and y points are both on a line, then we can use homographies to fix these lines, and assume that xi = (ai, 0, 1) and yi = (bi, 0, 1) for i = 1, …, 4. This means that we can identify xi with (ai, 1) ∈ ℙ1 and yi with (bi, 1) ∈ ℙ1. Under this substitution, Z4 has exactly one non-zero minor

det(M)=deta1b1a1b11a2b2a2b21a3b3a3b31a4b4a4b41=[14]x[23]x[13]y[24]y[13]x[24]x[14]y[23]y. (18)

The second equality in (18) can be checked in Macaulay2. The bracket expression set to 0 is the equality of cross-ratios

(x1,x2;x3,x4)=(y1,y2;y3,y4)

which becomes well-defined when the xi’s are distinct and the yi’s are distinct.□

5 The case k = 5

We now consider the rank deficiency of Z5. In this case, we present two theorems. Theorem 5.1 is a characterization of rank deficiency in terms of brackets. We will see that a necessary condition for rank drop is that one set of points (say {yi}) is on a line . Theorem 5.2 gives a geometric characterization of rank deficiency when the points are sufficiently generic, i.e., when the yi’s on the line are distinct and the xi’s are in general position. In this case the rank drop of Z5 is equivalent to the existence of an isomorphism from the conic ω through the xi’s to the line taking xi to yi. This is analogous to Theorem 4.2 (3)(b) where there is a homography taking xi to yi under sufficient genericity. We will illustrate both results using Example 2.

Theorem 5.1

Given a configuration {(xi,yi)}i=15 in2 × ℙ2, the matrix Z5 is rank deficient if and only if either one of the inherited conditions holds, or the points in one2 (say in Py2 ) are on a line and the following bracket equations hold for all distinct i1, i2, i3, i4, j ∈ {1, 2, 3, 4, 5}:

[i1i3j]x[i2i4j]x[i1i4]y[i2i3]y=[i1i4j]x[i2i3j]x[i1i3]y[i2i4]y. (19)

By the action of permutations, there are only 5 equations to consider in (19), one for each choice of j ∈ {1, 2, 3, 4, 5}. If both the xis and the yis are contained in lines then (19) holds automatically. Generically, (19) is equivalent to the cross-ratio equalities

(xi1,xi2;xi3,xi4;xj)=(yi1,yi2;yi3,yi4) (20)

for all distinct i1, i2, i3, i4, j ∈ {1, 2, 3, 4, 5}.

Proof

The proof splits into two cases. For the first case, we assume that the points in neither side are entirely on a line and show that rank will drop only under inherited conditions. For the second case, we are then free to assume that one side is contained in a line, which allows us to simplify the Macaulay2 calculations.

  1. Suppose that neither the xi’s nor the yi’s are on a line, and none of the inherited conditions hold. Then we can assume without loss of generality that x1x2 and y1y2 because no side has four coincident points (due to the absence of inherited conditions). Since neither side is contained entirely in a line, we can also assume that x1, x2, x3 are not collinear. Then either y1, y2, y3 are also not collinear, or y1, y2, y3 are on a line which means that we can assume that y1, y2, y4 are not collinear. Applying homographies, we therefore need to work with two different fixings of points:

    x1=(1,0,0)=y1 and x2=(0,1,0)=y2 and x3=(0,0,1)=y3 (21)

    and

    x1=(1,0,0)=y1 and x2=(0,1,0)=y2 and x3=(0,0,1)=y4. (22)

    Note that these fixings are not to finite points as we have done so far, but are equivalent to the fixing in (4) as remarked after Corollary 2.2. They allow for faster computations in Macaulay2.

    In (21), the ideal generated by the 5 × 5 minors of Z5 is the intersection of 20 prime ideals. Of these, 4 correspond to invalid inputs xi, yi = (0, 0, 0); 7 correspond to the inherited repeated point condition from Theorem 4.2(1); 6 correspond to the inherited condition from Theorem 4.2(2); and the remaining 3 correspond to the inherited bracket/cross-ratio condition from Theorem 4.2(3)(b).

    In (22), the ideal generated by the 5 × 5 minors of Z5 is the intersection of 17 prime ideals. Of these, 4 correspond to invalid inputs xi, yi = (0, 0, 0); 7 correspond to repeated points; 4 correspond to the inherited condition from Theorem 4.2(2); and the remaining 2 correspond to the inherited bracket/cross-ratio condition from Theorem 4.2(3)(b).

    Thus, there are no new conditions for rank drop if neither the xi’s nor the yi’s are in a line.

  2. Suppose that y1, …, y5 are on a line. Applying homographies, we may assume that yi = (mi, 0, 1) for i = 1, …, 5 and all xi are finite. We may further assume that x1 = (0, 0, 1) and that y1 = (0, 0, 1). Then rearranging the columns of Z5 we get

    Z5=000001000x21m2x22m2x21x22m21000x31m3x32m3x31x32m31000x41m4x42m4x41x42m41000x51m5x52m5x51x52m51000.

    Let I be the ideal generated by the 6 non-zero 5 × 5 minors of Z5 and let J be the ideal generated by the 5 differences of left and right hand sides in the bracket equations (19) under the same fixings of points. Then we can verify using Macaulay2 that I = J . A symmetric argument will switch the roles of the x and y points.□

Example 2

The following example illustrates Theorem 5.1 and motivates Theorem 5.2. Consider

x1=(0,0,1)y1=(0,0,1)x4=(1,1,1)y4=(4,0,1)x2=(1,0,1)y2=(1,0,1)x5=(50,98,113)y5=(8,0,1)x3=(0,1,1)y3=(3,0,1)

Here the yi’s are on the line yi = (mi, 0, 1) and we can verify that (19) is satisfied. The matrix

Z5=0000000011010001010003013014014014014000507840989040113

has rank 4.

Note that in this example, the yi’s (which are on a line) are distinct, and that no 3 of the xi’s are on a line. In this case, there is a 1-parameter family of rank 2 projective transformations {Tλ} such that Tλ (xi) = yi for i = 1, …, 5 except for 5 unique values of λ, one for each of x1, …, x5. In this example, this is the family

Tλ=0120737+λ7150750. (23)

For each i = 1, …, 5 there exists unique λi such that Tλi has nullvector (center) xi. These are

λ1=λ2=0λ3=45λ4=32λ5=2120. (24)

Since Tλi (xi) = (0, 0, 0) ≁ yi, these 5 transformations can be seen as degenerate members of the family.

We now formalize the second half of Example 2 as a theorem. As seen in Theorem 5.1, for Z5 to drop rank, either the xi’s or the yi’s have to be on a line. Suppose the yi’s are distinct and on a line, and the xi’s are in general position. Then the xi’s lie on a conic ω and we will see that there is an isomorphism taking ω to the line containing the yi’s so that each xi maps to yi.

Theorem 5.2

Suppose that we have a configuration {(xi,yi)}i=15 such that no 3 of the xis are collinear and the yis are distinct and on a line y. Let ω be the conic through x1, …, x5. Suppose further that no inherited condition for rank drop holds. Then the following are equivalent:

  1. The matrix Z5 is rank deficient.

  2. For all cω ∖ {x1, …, x5} there exists a projective transformation Tc : ℙ2 → ℙ1 centered at c such that Tc (xi) ∼ yi for all i.

  3. For some cω ∖ {x1, …, x5} there exists a projective transformation Tc : ℙ2 → ℙ1 centered at c such that Tc(xi) ∼ yi for all i.

  4. There exists a cross-ratio preserving isomorphism F : ωy such that F(xi) = yi for all i.

Proof

To prove that (1) implies (2), let cω ∖ {x1, …, x5}. Then for all distinct i1, i2, i3, i4, j ∈ {1, …, 5}

(yi1,yi2;yi3,yi4)=(xi1,xi2;xi3,xi4;xj)=(xi1,xi2;xi3,xi4)ω=(xi1,xi2;xi3,xi4;c) (25)

for any cω. The first equation is from (20) in Theorem 5.1 and the other two equalities follow by Definition 3.6. Let cxi be the line through c and xi and let be any line such that c. Then if we define ai to be the unique intersection of and cxi it follows that

(ai1,ai2;ai3,ai4)=(xi1,xi2;xi3,xi4;c)=(yi1,yi2;yi3,yi4) (26)

for all i1, i2, i3, i4 ∈ {1, …, 5}. See Figure 2. It follows that there exists homography H such that Haiyi for all i. Therefore if Tc is the projection in Px2 centered at c onto the line , then Tc := H Tc is the desired projective transformation with Tc(xi) ∼ yi for all i.

Figure 2 
We construct the projection with center c.
Figure 2

We construct the projection with center c.

Clearly (2) implies (3). To show that (3) implies (1), suppose that such a projection exists and is centered at c. Then

(xi1,xi2;xi3,xi4;xj)=(xi1,xi2;xi3,xi4)ω=(xi1,xi2;xi3,xi4;c)=(yi1,yi2;yi3,yi4) (27)

for all distinct i1, i2, i3, i4, j ∈ {1, …, 5}. The first two equalities follow by Definition 3.6 and the last equality follows by hypothesis. It immediately follows that Z5 is rank deficient by Theorem 5.1.

To show that (1) implies (4) we may assume that the cross-ratio equalities (20) hold. Then define an isomorphism F : ωy by F(x1) = y1, F(x2) = y2, F(x3) = y3. Then for all pω ∖ {x1, x2, x3}, F(p) = q is such that (x1, x2; x3, p)ω = (y1, y2; y3, q). It then follows that F(x4) = y4 and F(x5) = y5 because

(x1,x2;x3,x4)ω=(y1,y2;y3,y4)(x1,x2;x3,x5)ω=(y1,y2,y3;y5) (28)

by hypothesis (Theorem 5.1).

Finally, to show that (4) implies (1), suppose that such an isomorphism F exists. Then

(xi1,xi2;xi3,xi4;xj)=(xi1,xi2;xi3,xi4)ω=(F(xi1),F(xi2);F(xi3),F(xi4))=(yi1,yi2;yi3,yi4) (29)

for all i1, i2, i3, i4, j ∈ {1, …, 5}. The first equality follows by Definition 3.6 and the second and third equalities follow by hypothesis. It then follows that Z5 is rank deficient by Theorem 5.1.□

6 The case k = 6

The final case we will consider in this paper is a set of 6 inputs {(xi,yi)}i=16P2×P2. We organize our results in two parts. In the first part we state algebraic characterizations of the rank deficiency of Z6. These are analogous to the results in the previous sections in that the results are in terms of invariant theory. In the second part, we explain the geometry that leads to these algebraic statements. This relies on the classical theory of cubic surfaces and their combinatorics. The geometry works in an open region where the input is sufficiently generic. However, as we saw in previous sections, the algebra that it translates to describes the Zariski closure of these open regions and hence the set of all inputs that make Z6 rank deficient.

We say that p1, …, p6 ∈ ℙ2 are in general position if at most 2 points are on a line and at most 5 points are on a conic. In several results below we need that the x points and/or the y points are in general position. We will also need that a collection of point pairs {(xi, yi)} ⊂ ℙ2 × ℙ2 is in general position, by which we mean that

  1. the x points and y points are in general position, and

  2. no more than 4 pairs (xi, yi) are related by a homography Hxi = yi.

Note that if (1) holds then Z5 has full rank because all of our previous conditions for rank drop require at least the xi’s or the yi’s to be on a line.

The variety ℙ2 × ℙ2 can be embedded in ℙ8 as the image of the Segre map

P2×P2P8 such that (x,y)xy. (30)

In what follows it will be helpful to identify ℙ8 with the set of all complex 3 × 3 matrices up to scaling via the bijection M ↔ vec(M). Then the Segre embedding of ℙ2 × ℙ2 in ℙ8 (which we also call ℙ2 × ℙ2) can be identified with the set of rank one 3 × 3 matrices up to scaling since

xy=(vec(yx)).

The Segre variety ℙ2 × ℙ2 ⊂ ℙ8 has dimension 4 and degree 6. It is the zero locus of the determinantal ideal generated by the 2 × 2 minors of a symbolic 3 × 3 matrix whose entries denote the coordinates of ℙ8.

We first show that given 5 points {(xi,yi)}i=15 in general position, there exists a unique 6th point (x6, y6) ∈ ℙ2 × ℙ2 such that {(xi,yi)}i=16 is rank deficient.

Lemma 6.1

Let {(xi,yi)}i=15 be in general position. Then there is a unique new 6th point (x6, y6) ∈ ℙ2 × ℙ2 such that the matrix Z6 is rank deficient. This 6th point is a rational function of the first 5; if we use homographies to fix the first 4 points as

x1=(1,0,0)=y1x2=(0,1,0)=y2x3=(0,0,1)=y3x4=(1,1,1)=y4,

then the 6th point has the formula

x6=(y53y52y53x52x53y52,y53y51y53x51x53y51,y51y52y51x52x51y52)y6=(x53x52y53x52x53y52,x53x51y53x51x53y51,x51x52y51x52x51y52). (31)

Proof

Let 𝓡(Z5) denote the row span of Z5 = (xiyi)i=15 . By general position, rank(Z5) = 5 which means that 𝓡(Z5) ≅ ℙ4. Therefore, generically, 𝓡(Z5) intersects the 4-dimensional degree 6 Segre variety ℙ2 × ℙ2 transversally in 6 distinct points. Each of these points corresponds to a rank one matrix yx which we have identified with xy. Five of these intersection points are the matrices xiyi for i = 1, …, 5 and the 6th point x6y6 corresponds to a unique point (x6, y6) ∈ ℙ2 × ℙ2.

To prove the second part of the lemma, we first use homographies to fix the first 4 points as mentioned. We can then use Macaulay2 to verify that we can solve for (x6, y6) in terms of (x5, y5) as in equations (31).□

The existence of, and constructions for, this 6th point were noted in both [25] and [31], but not the rational formula. The rational formula shows that if (xi,yi)i=15 are real then (x6, y6) will be real as well. Note that if {xi}i=15 and {yi}i=15 were related by homography then the denominators in (31) would all be 0. Here is an illustration of Lemma 6.1.

Example 3

We begin with 5 point pairs

x1=(1,0,0)y1=(1,0,0)x2=(0,1,0)y2=(0,1,0)x3=(0,0,1)y3=(0,0,1)x4=(1,1,1)y4=(1,1,1)x5=(3,5,1)y5=(8,2,1).

If we leave (x6, y6) symbolic and construct the matrix Z6, we can verify with Macaulay2 that the unique new point pair that will result in rank deficiency is x6(13,75,317),y6(43,25,117), which is exactly what we get from the formula above.

In Lemma 6.16 we will prove a counterpart to the above result, where we instead begin with x1, …, x6 and calculate (up to homography) the points y1, …, y6 that will result in rank deficiency.

6.1 Algebraic characterizations of rank deficiency

In this subsection we will write down various algebraic characterizations of rank deficiency. While these statements are by no means obvious or intuitive, once stated, they are straightforward to check using Macaulay2. In the next section we explain the geometric origins of these statements.

Our first result is a characterization of rank drop using brackets and cross-ratios as in the previous sections.

Definition 6.2

Given 6 input points (xi, yi) ∈ ℙ2 × ℙ2, let 𝓑rackets(Z6) be the ideal generated by the differences of the two sides in the following bracket equations

[ikp]x[jrp]x[irs]y[jks]y=[irp]x[jkp]x[iks]y[jrs]y (32)

for all distinct indices i, j, k, r, p, s ∈ {1, 2, 3, 4, 5, 6}.

Up to permutations, there are 30 equalities in (32): one for each choice of p and s. Under sufficient genericity, the bracket equations are the planar cross-ratio equalities

(xi,xj;xk,xr;xp)=(yi,yj;yk,yr;ys). (33)

Next we define a specific kind of highly degenerate configuration.

Definition 6.3

We say that {(xi,yi)}i=16 is an asymmetric double triangle if the {xi} and {yi} each consist of 3 distinct points such that, up to reordering, y1 = y2, y3 = y4, y5 = y6, x1 = x4, x2 = x5, and x3 = x6.

Lemma 6.4

Let {(xi,yi)}i=16 be an asymmetric double triangle. Then Z6 is rank deficient if and only if an inherited condition holds.

Proof

We may assume up to reordering that y1 = y2 =: v1, y3 = y4 =: v2, y5 = y6 =: v3, x1 = x4 =: u1, x2 = x5 =: u2, and x3 = x6 =: u3. If u1, u2, u3 and v1, v2, v3 are both non-collinear, then we can assume by change of coordinates that u1 = v1 = (1, 0, 0), u2 = v2 = (0, 1, 0) and u3 = v3 = (0, 0, 1). It is then simple to check that Z6 has full rank.

If the three of points of one type are distinct but collinear and the three points of the other type are non-collinear, then we may assume up to change of coordinates and relabeling that u1 = v1 = (1, 0, 0), u2 = v2 = (0, 1, 0), u3 = (1, 1, 0), and v3 = (0, 0, 1). Again, it is simple to check that Z6 has full rank.

If two points of one type are coincident, say u1 = u2, then an inherited condition holds and the rank drops. If u1, u2, u3 and v1, v2, v3 are both collinear then again the rank drops due to an inherited condition.□

Let 𝓜inors(Z6) denote the ideal of maximal minors of Z6. The following statement can be checked using Macaulay2.

Theorem 6.5

Let {(xi,yi)}i=16 be such that neither the xis nor the yis are on a line and {(xi,yi)}i=16 is not an asymmetric double triangle. Then Z6 is rank deficient if and only if either one of the inherited conditions holds or the 30 bracket equations (32) hold.

Proof

As in the proof of Theorem 5.1 we split into two cases. Applying homographies we can assume that

x1=(1,0,0)=y1x2=(0,1,0)=y2x3=(0,0,1)=y3

or

x1=(1,0,0)=y1x2=(0,1,0)=y2x3=(0,0,1)=y4.

In the first case the ideal 𝓜inors(Z6) decomposes into 19 components. The first 6 correspond to invalid inputs xi, yi = (0, 0, 0), the next 12 correspond to repeated point pairs, and the 19th component we refer to as M19. The ideal 𝓑rackets(Z6) decomposes into 13 components. One component is M19 and the other 12 components correspond to asymmetric double triangles.

In the second case the ideal 𝓜inors(Z6) decomposes into 17 components. The first 6 correspond to invalid inputs xi, yi = (0, 0, 0), the next 10 correspond to repeated point pairs, and the 17th component we refer to as M17. The ideal 𝓑rackets(Z6) decomposes into 21 components. One component is M17 and the other 20 components correspond to asymmetric double triangles.□

In Theorem 6.5 the only assumption is that neither the xi’s nor the yi’s are on a line. In Theorem 6.8 we will tackle the case of one side in a line, and together the two theorems cover all possibilities.

Before we can do that, we consider an important subcase of Theorem 6.5, namely the generic case. The main result here is Theorem 6.7 which uses tools from invariant theory that we have not encountered thus far. Besides being a key result of this paper, it will also lead to a proof of Theorem 6.8, completing the story.

Suppose that x1, …, x6 Px2 and y1, …, y6 Py2 are in general position. Then we may fix

x1=y1=(1,0,0),x2=y2=(0,1,0),x3=y3=(0,0,1),x4=y4=(1,1,1) (34)

and consider the rank deficiency of Z6 whose 5th and 6th rows are left symbolic as

(x51,x52,x53)(y51,y52,y53),(x61,x62,x63)(y61,y62,y63). (35)

Recall that the general position assumption guarantees that no submatrix of Z6 with 5 rows is rank deficient. Under these assumptions, one can use Macaulay2 to check that 𝓜inors(Z6) is a radical ideal of dimension 9 and degree 4. Its prime decomposition consists of 14 components. The first 4 of these components correspond to invalid inputs xi, yi = (0, 0, 0), the next 9 correspond to repeated points, and the 14th component, denoted as M14, encodes a new set of conditions, and has dimension 8 and degree 35. Since Z6 is rank deficient if and only if all its maximal minors vanish, we conclude that the new conditions for rank drop are exactly that (x5, y5) and (x6, y6) lie in the variety of M14.

We will now interpret M14 in terms of the invariants of x1, …, x6 and y1, …, y6. For a set of 6 points p1, …, p6 ∈ ℙ2 define

[(ij)(kl)(rs)]:=[ijr][kls][ijs][klr]. (36)

This is a classical invariant of points in ℙ2 whose vanishing expresses that the three lines pipj, pkpl and prps meet in a point; see [6, page 169]. Using these invariants, Coble defines the following 6 scalars in [6, page 170]:

a¯=[(25)(13)(46)]+[(51)(42)(36)]+[(14)(35)(26)]+[(43)(21)(56)]+[(32)(54)(16)]b¯=[(53)(12)(46)]+[(14)(23)(56)]+[(25)(34)(16)]+[(31)(45)(26)]+[(42)(51)(36)]c¯=[(53)(41)(26)]+[(34)(25)(16)]+[(42)(13)(56)]+[(21)(54)(36)]+[(15)(32)(46)]d¯=[(45)(31)(26)]+[(53)(24)(16)]+[(41)(25)(36)]+[(32)(15)(46)]+[(21)(43)(56)]e¯=[(31)(24)(56)]+[(12)(53)(46)]+[(25)(41)(36)]+[(54)(32)(16)]+[(43)(15)(26)]f¯=[(42)(35)(16)]+[(23)(14)(56)]+[(31)(52)(46)]+[(15)(43)(26)]+[(54)(21)(36)] (37)

Let āx, x, x, x, ēx, x be the polynomials obtained by computing the expressions in (37) using 6 symbolic vectors x1, …, x6 where each xi = (xi1, xi2, xi3). Similarly, let āy, y, y, y, ēy, y be obtained from symbolic vectors y1, …, y6 where each yi = (yi1, yi2, yi3).

Definition 6.6

Let 𝓘nvariants(Z6) denote the ideal of 2 × 2 minors of

a¯xb¯xc¯xd¯xe¯xf¯xa¯yb¯yc¯yd¯ye¯yf¯y (38)

after specializing the first 4 points (xi, yi) as in (34).

The ideal 𝓘nvariants(Z6) is radical of dimension 9 and degree 4. Its prime decomposition consists of 14 components. The first 4 of these components correspond to invalid inputs xi, yi = (0, 0, 0), the next 8 correspond to 3 repeated points in either Px2 or Py2 , the 13th component corresponds to x5y5 and x6y6 (which encodes that {xi} and {yi} are related by a homography). The 14th component, denoted as I14, provides a new condition. This ideal has dimension 8 and degree 35.

Theorem 6.7

Given a configuration {(xi,yi)}i=16 such that the {xi} and {yi} are in general position, one has

a¯xb¯xc¯xd¯xe¯xf¯xa¯yb¯yc¯yd¯ye¯yf¯y (39)

if and only if either the matrix Z6 is rank deficient or yi = Hxi for some homography H ∈ PGL(3).

Proof

It can be checked using Macaulay2 that M14 = I14.□

The statement of Theorem 6.7 begs the question of how homography ties in with rank drop, if at all. We give an example to show that the existence of a homography such that yi = Hxi for 6 points (xi, yi) is not sufficient for Z6 to be rank deficient. In contrast, it does suffice for 7 point pairs.

Example 4

By Lemma 2.1, if there exists H ∈ PGL(3) such that yi = Hxi for i = 1, …, 6, then we may assume that xi = yi for i = 1, …, 6. Then the null space of Z6 consists exactly of the vectors vec(F) for the 3 × 3 matrices F such that

xiFxi=0for i=1,,6. (40)

Generically, this is exactly the set of 3 × 3 skew-symmetric matrices, Skew3 ≅ ℙ2. Hence rank(Z6) = 6. For example, consider

x1=y1=(1,0,0)x2=y2=(0,1,0)x3=y3=(0,0,1)x4=y4=(1,1,1)x5=y5=(2,3,1)x6=y6=(3,7,1)

for which the nullspace of Z6 is Skew3. In general, if the homography is given by the matrix H then the null space of Z6 is obtained by left action of H on Skew3. We note that this null space is unusual in that will always consist entirely of rank 2 matrices.

To finish our algebraic discussion, we consider the case of one ℙ2 having all input points in a line. In order to state the result, we introduce the classical Joubert invariants of 6 points in ℙ1, compare [6], which are

Ay=[25][13][46]+[51][42][36]+[14][35][26]+[43][21][56]+[32][54][16]By=[53][12][46]+[14][23][56]+[25][34][16]+[31][45][26]+[42][51][36]Cy=[53][41][26]+[34][25][16]+[42][13][56]+[21][54][36]+[15][32][46]Dy=[45][31][26]+[53][24][16]+[41][25][36]+[32][15][46]+[21][43][56]Ey=[31][24][56]+[12][53][46]+[25][41][36]+[54][32][16]+[43][15][26]Fy=[42][35][16]+[23][14][56]+[31][52][46]+[15][43][26]+[54][21][36]. (41)

Theorem 6.8

Suppose that y1, …, y6 Py2 are on a line. Then Z6 is rank deficient if and only if

a¯xAy+b¯xBy+c¯xCy+d¯xDy+e¯xEy+f¯xFy=0. (42)

Furthermore, if x1, …, x6 are in general position and y1, …, y6 are distinct, then (42) holds if and only if there exists a projection T : Px2 Py1 such that T(xi) ∼ yi for all i.

This theorem is analogous to Theorems 4.2 (3)(b) and 5.2. The proof of this theorem will follow from the geometry that we develop in the next subsection.

6.2 Geometric characterizations of rank deficiency

We next explain where the algebraic results (and ideals) from the previous section come from by providing a geometric characterization of rank deficiency of Z6. This relies on the classical theory of smooth cubic surfaces in ℙ3. For the convenience of the reader, we have collected the needed facts about cubic surfaces in Appendix A.

6.2.1 Cubic surfaces and Schäfli double sixes

A cubic surface in ℙ3 is a hypersurface defined by a single homogenous polynomial of degree 3 in ℂ[z0, z1, z2, z3] where (z0, …, z3) are the coordinates of ℙ3.

Definition 6.9

A pair of 6 lines {1, …, 6} and {1,,6} in ℙ3 is called a Schläfli double six if

  1. the 6 lines in each set are mutually skew (i.e., no 2 of them intersect), and

  2. i intersects j if and only if ij.

A smooth cubic surface has 36 Schläfli double sixes. The rich and highly symmetric properties of these line configurations are listed in Appendix A and lie at the heart of all of our major results in this section. Our main (geometric) theorem in this section is the following.

Theorem 6.10

Suppose that {(xi,yi)}i=16 ⊂ ℙ2 × ℙ2 is in general position. Then Z6 is rank deficient if and only if there exists a cubic surface 𝓢 that can be obtained as the blow up of both Px2 in x1, …, x6 and Py2 in y1, …, y6 such that their respective exceptional curves {1, …, 6} and {1,,6} form a Schläfli double six on 𝓢.

When Z6 is rank deficient, this cubic surface has the determinantal representation 𝓝(Z6) ∩ 𝓓, where 𝓝(Z6) denotes the right null space of Z6 and 𝓓 is the cubic hypersurface of all rank deficient 3 × 3 matrices.

To prepare for the proof of this theorem, we begin by constructing a cubic surface from 5 points {(xi,yi)}i=15 in general position. Recall that we can identify a point in ℙ8 with a 3 × 3 matrix M via its vectorization vec(M). Under this identification, 𝓓 ⊂ ℙ8 is the cubic hypersurface 𝓓 = {M : det(M) = 0}. Let 𝓝(Z5) ⊂ ℙ8 be the null space of Z5. Since {(xi,yi)}i=15 is in general position, rank(Z5) = 5 and so 𝓝(Z5) ≅ ℙ3. Therefore

S:=N(Z5)D (43)

is a cubic surface in 𝓝(Z5) ≅ ℙ3 ⊂ ℙ8. This surface is smooth by the general position assumption. (We note that if {xi} and {yi} were related by a homography then S would be degenerate; see Example 4.)

The surface S has a natural determinantal representation (see Appendix A for the definition): Pick a basis M0, …, M3 of 𝓝(Z5) and set

M(z)=z0M0++z3M3 (44)

where zi ∈ ℂ. If (xi, yi) with i = 1, …, 5 are real, we can choose M0, …, M3 to be real matrices. Since M(z) is a parameterization of 𝓝(Z5), S is cut out by det(M(z)) = 0, and M(z) is a determinantal representation of S. Generically, the 3-dimensional 𝓝(Z5) does not intersect the 4-dimensional ℙ2 × ℙ2 of rank one matrices, and so a point M(z) on S has rank exactly 2. This also follows from S being smooth.

Next we construct a Schläfli double six on the cubic surface S = 𝓝(Z5) ∩ 𝓓 from {(xi,yi)}i=15 .

Definition 6.11

Given {(xi,yi)}i=15 in general position, define

xi:={M(z):M(z)xi=0} for i=1,,5, and  (45)
yj:={M(z):yjM(z)=0} for j=1,,5, (46)

where M(z) is of the form (44).

For each i = 1, …, 5 we have xi ⊂ 𝓝(Z5) by the definition of M(z) in (44). Further, since M(z) xi = 0, M(z) has a non-trivial right null space and hence det(M(z)) = 0. Therefore xiS = 𝓝(Z5) ∩ 𝓓. The condition that M(z) ∈ 𝓝(Z5) is equivalent to M(z) satisfying the 5 linear constraints

xiyi,M(z)=yiM(z)xi=0 for i=1,,5. (47)

Therefore, xi is cut out by 8 linear constraints including the 3 constraints given by M(z) xi = 0. However one of them is redundant since M(z) xi = 0 implies that yiM(z)xi=0. Therefore, xi is cut out by 7 independent linear constraints and is hence a line on the cubic surface S. Similarly, each yj with j = 1, …, 5 is a line on S.

Lemma 6.12

The lines xi and yj satisfying the following incidence relations:

  1. The lines {xi, i = 1, …, 5} are mutually skew.

  2. The lines { yj , j = 1, …, 5} are mutually skew.

  3. xi intersects yj if and only if ij.

Therefore, {xi, i = 1, …, 5} and { yj , j = 1, …, 5} form part of a unique Schläfli double six on S.

Proof

The proof of the incidence relations follows from dimension counting and can be found in [25, Lemma 6.1]. The two sets of lines are part of a unique Schläfli double six on S since each double six is uniquely defined by three of its line pairs, and we have 5 line pairs here with the correct incidences; see Lemma A.1 (3).□

By Lemma A.1, given a Schläfli double six on S, there are two sets of 6 points {pi} ⊂ Pp2 and {qj} ⊂ Pq2 such that S is the blow up of Pp2 at {pi} with blow up morphism πp : S Pp2 , and also the blow up of Pq2 at {qj} with blow up morphism πq : S Pq2 . Each set of skew lines in the given double six is the set of exceptional curves of these point sets under πp and πq . The blow up morphism πp : S Pp2 sends M(z) ∈ S to the unique p ∈ ℙ2 in its right null space and the blow up morphism πq : S Pq2 sends M(z) ∈ S to the unique q ∈ ℙ2 in its left null space. See the second half of Appendix A.

We are now ready to prove the forward direction of Theorem 6.10.

Definition 6.13

Let 6, 6 denote the 6th pair of lines in the Schläfli double six on S consisting of {xi}i=15 and {yj}j=15 from Lemma 6.12.

Lemma 6.14

Suppose that {(xi,yi)}i=15 is in general position. Then the addition of a new row x6y6 makes Z6 rank deficient if and only if {xi}i=16and{yj}j=16 form a Schläfli double six on S where

6=x6={M(z):M(z)x6=0}and6=y6={M(z):y6M(z)=0}. (48)

Proof

To prove the forward direction, suppose Z6 is rank deficient. Then 𝓝(Z5) = 𝓝(Z6) and S = 𝓝(Z6) ∩ 𝓓. Check that x6 is a line on S by the same argument as for xi with i ≤ 5. Similarly, y6 is a line on S. One can also check that the collection {xi}i=16and{yj}j=16 satisfies the incidence relations of a Schläfli double six. This must mean that 6 = x6 and 6 = y6 since the first 5 line pairs determine a unique double six on S.

To prove the reverse direction, suppose we have a Schläfli double six on S as in the statement of the lemma. Then by the discussion before Definition 6.13, S is the blow up of Px2 at {xi} under a morphism πx and S is also the blow up of Py2 at {yi} under a morphism πy . Further, πx(xi) = xi for i = 1, …, 6 and πy ( yj ) = yj for j = 1, …, 6. We need to argue that Z6=(xiyi)i=16 is rank deficient. Recall Lemma 6.1 and let (x0, y0) be the unique new point pair such that Z′ with rows xiyi for i = 0, …, 5 is rank deficient. Then, by the forward direction, we must have 6 be the exceptional curve of x0 under πx and hence x0 = πx(6) = x6. Similarly, y0 = πy( 6 ) = y6. We conclude that (x0, y0) = (x6, y6) and therefore Z6 is rank deficient.□

The following corollary to Lemma 6.14 and Theorem A.5 concludes the proof of the forward direction of Theorem 6.10.

Corollary 6.15

If {(xi,yi)}i=16 is in general position and Z6 is rank deficient, then the cubic surface S = 𝓝(Z6) ∩ 𝓓 is the blow up of Px2 at x1, …, x6 with exceptional curves xi, i = 1, …, 6, and also the blow up of Py2 at y1, …, y6 with exceptional curves yj , j = 1, …, 6. The two sets of lines form a Schläfli double six on S.

Note that Lemma 6.14 gives a second construction for the unique point pair (x6, y6) in Lemma 6.1, namely as the images under the blow up morphisms of the final pair of lines in the unique Schläfli double six on S determined by the lines {xi}i=15 and {yj}j=15 . This was also noted in [25].

To prove the backward direction of Theorem 6.10, we will first prove a counterpart to Lemma 6.1.

Lemma 6.16

If x1, …, x6 are in general position, then there is a unique set of 6 points y1, …, y6 in general position (up to homography) such that Z6 is rank deficient. If we assume that

x1=(1,0,0)=y1x2=(0,1,0)=y2x3=(0,0,1)=y3x4=(1,1,1)=y4

then we can obtain y5, y6 via the rational function

y5=(x63x62x53x62x63x52,x63x61x53x61x63x51,x62x61x61x52x51x62) (49)
y6=(x53x52x53x62x63x52,x53x51x53x61x63x51,x52x51x61x52x51x62). (50)

Proof

We first prove that this construction will result in rank deficiency; the uniqueness will come afterwards as a consequence of Corollary 6.15. We construct the matrix Z6=(xiyi)i=16 and fix x1, …, x4, y1, …, y6 as in the statement of the lemma. We can then verify with Macaulay2 that all 6 × 6 minors of Z6 are 0 and therefore Z6 is rank deficient.

To see that this choice is unique, construct S = 𝓝(Z6) ∩ 𝓓. Then by Corollary 6.15, S is both the blow up of Px2 at x1, …, x6 and the blow up of Py2 at y1, …, y6 and the two sets of exceptional curves {x1, …, x6}, {y1,,y6} form a Schläfli double six. Now suppose there was some other y1,,y6 in general position such that Z6 = (xiyi)i=16 was rank deficient. Then S′ = 𝓝( Z6 ) ∩ 𝓓 is also the blow up of Px2 at x1, …, x6. Moreover, it is also the blow up of Py2 at y1,,y6 and the exceptional curves x1,,x6,y1,,y6 form a Schläfli double six. It follows by Lemma A.2 that yi=yi for all i. If we compose the blow up morphisms πyπy1:Py2Py2 we obtain an invertible projective transformation; thus this composition is a homography taking yi yi . This concludes the proof.□

The formula shows that if x1, …, x6 are real then, up to homography, y1, …, y6 will also be real.

Example 5

Consider Example 3 again and now fix the points

x1=(1,0,0)x2=(0,1,0)y1=(1,0,0)y2=(0,1,0)x3=(0,0,1)x4=(1,1,1)y3=(0,0,1)y4=(1,1,1)x5=(3,5,1)x6=(13,75,317).

If we leave y5, y6 symbolic and construct the matrix Z6, we can verify with Macaulay2 that the unique new points that will result in rank deficiency are y5 ∼ (8, 2, 1), y6(43,25,117), which is exactly what we get from the formula above. Moreover, check that this matches up exactly with the points in Example 3.

The following corollary now completes the reverse direction of Theorem 6.10.

Corollary 6.17

Given a smooth cubic surface 𝓢 with a Schläfli double six {i}i=16and{j}i=16 that arises as the exceptional lines of x1, …, x6 Px2 and y1, … y6 Py2 , the matrix Z6 = (xiyi)i=16 is rank deficient.

Proof

Lemma 6.16 provides %%us with unique (up to homography) points y1,,y6 Py2 such that Z6=(xiyi)i=16 is rank deficient. Then, following along the same lines as the proof of Lemma 6.16, S = 𝓝( Z6 ) ∩ 𝓓 and the points {yi} and {yi} must be related by homography. It follows by Lemma 2.1 that Z6 is rank deficient.□

Combining Corollaries 6.15 and 6.17 we obtain a method to compute a determinantal representation of a smooth cubic surface starting with a Schläfli double six on it.

Corollary 6.18

Let 𝓢 be a smooth cubic surface with a Schläfli double six {i} and {j}, and let {(xi,yi)}i=16P2×P2 be such that the i are the exceptional lines of the blow up of Px2 at x1, …, x6 and 6 are the exceptional lines of the blow up of Py2 at y1, … y6. Then a determinantal representation of 𝓢 is given by M(z) = ∑ ziMi where M0, …, M3 is a basis of the null space of Z6=(xiyi)i=16.

This gives us the final statement of Theorem 6.10, concluding the proof.

Remark 6.19

In the classical literature one finds a determinantal representation of a cubic surface from 6 points in a ℙ2 (whose blow up at these points is the surface), via the Hilbert–Burch theorem (see Appendix A). It is worthwhile to contrast that method with that in Corollary 6.18 where the starting point are 6 point pairs in ℙ2 × ℙ2. Then a determinantal representation (44) of the surface can be obtained from a basis of 𝓝(Z6) using simple linear algebra.

6.2.2 Conic intersections and the bracket ideal 𝓑rackets(Z6)

We now use Theorem 6.10 to explain the ideal 𝓑rackets(Z6) used in the algebraic characterization of rank deficiency in Theorem 6.5.

Consider a Schläfli double six {i}i=16,{j}j=16 on a cubic surface 𝓢 where the lines i are the exceptional lines under the blow up πx of points xi Px2 and the lines j are the exceptional lines under the blow up πy of points yj Py2 . Let Ci be the unique conic passing through {x1, …, x6}∖{xi} and Ci the unique conic passing through {y1, …, y6}∖{yi}. By Lemma A.1, the lines i are the strict transforms of the conics Ci Py2 and and the lines j are the strict transforms of the conics Cj Px2 . There is then a birational automorphism Px2 S Py2 such that xi Ci and Ciyi. This map is a degree 5 Cremona transformation and can be formalized using the following recipe due to Sturm [28], mentioned by Werner in [31]. These conics can be used to give a third construction of the unique 6th point pair (x6, y6) that can be added to {(xi,yi)}i=15 in general position so that Z6 is rank deficient. We state the result without proof in Lemma 6.20 and illustrate it in Figure 3. The data for this figure was chosen to make the conics easily visible.

Figure 3 
The point y6 is the unique intersection point of the 5 conics 
Ci′
$\begin{array}{}
\displaystyle
C'_i
\end{array}$ in 
Py2
$\begin{array}{}
\displaystyle
\mathbb{P}^2_y
\end{array}$.
Figure 3

The point y6 is the unique intersection point of the 5 conics Ci in Py2 .

Lemma 6.20

(See [28] and [31]). Let C be the unique conic through x1, …, x5 in general position, and let Hi be the unique homography such that yj = Hixj for all ji. Define Ci := Hi(C) for i = 1, …, 5. Then y6i=15Ci is the unique point in the intersection of the conics Ci .

Similarly, let Cbe the unique conic through y1, …, y5 and let Gj be the unique homography such that xi = Gjyi for all ij. Define Cj := Hj(C′) for j = 1, …, 5. Then x6j=15Cj is the unique point in the intersection of the conics Cj.

We can now restate Theorem 6.10 using Lemma 6.20 to yield another characterization of minimal rank deficiency of Z6.

Theorem 6.21

Let {(xi,yi)}i=15 be in general position. Then (x6, y6) is the unique new point under which Z6 is rank deficient if and only if y6 is the unique intersection point j=15Ci and x6 is the unique intersection point j=15Cj .

Theorem 6.21 explains the bracket and cross-ratio conditions that appear in Theorem 6.5. Let C be the conic through x1, …, x5 and let H be the homography such that Hxi = yi for i = 1, 2, 3, 4. Then C′ := HC is a conic that passes through y1 = Hx1, y2 = Hx2, y3 = Hx3, y4 = Hx4 and Hx5. By Lemma 3.5, y6C′ if and only if

det(H)[135]xdet(H)[245]x[146]y[236]y=[136]y[246]ydet(H)[145]xdet(H)[235]x. (51)

This can be rewritten as the bracket equation (32) by eliminating det(H):

[135]x[245]x[146]y[236]y=[136]y[246]y[145]x[235]x (52)

The other 29 bracket equations follow by selecting different conics and homographies.

6.2.3 The Cremona hexahedral form of a cubic surface

We now give a third characterization of minimal rank deficiency of Z6 via the Cremona hexahedral form of a cubic surface. This will explain the statement of Theorem 6.7.

Recall that every smooth cubic surface is the blow up of ℙ2 at 6 points in general position. The Cremona hexahedral form of the surface provides explicit equations for it in terms of the points being blown up. We briefly describe these equations and the properties we need. More details can be found in [6, Section 4], [13, Section 9.4.3] and [24, Section 5.6].

Given 6 points p1, …, p6 ∈ ℙ2, consider the following list of 6 cubic polynomials in u = (u0, u1, u2), the coordinates of ℙ2, that vanish on them:

a(u)=[25u][13u][46u]+[51u][42u][36u]+[14u][35u][26u]+[43u][21u][56u]+[32u][54u][16u]b(u)=[53u][12u][46u]+[14u][23u][56u]+[25u][34u][16u]+[31u][45u][26u]+[42u][51u][36u]c(u)=[53u][41u][26u]+[34u][25u][16u]+[42u][13u][56u]+[21u][54u][36u]+[15u][32u][46u]d(u)=[45u][31u][26u]+[53u][24u][16u]+[41u][25u][36u]+[32u][15u][46u]+[21u][43u][56u]e(u)=[31u][24u][56u]+[12u][53u][46u]+[25u][41u][36u]+[54u][32u][16u]+[43u][15u][26u]f(u)=[42u][35u][16u]+[23u][14u][56u]+[31u][52u][46u]+[15u][43u][26u]+[54u][21u][36u]. (53)

In fact, each summand in each cubic is the product of the lines through 3 disjoint pairs of the 6 points and already vanishes on the points. Formally these expressions are similar to the scalars in (37). These cubic polynomials are covariants of p1, …, p6 under the action of PGL(3), and appear in the work of Coble [6].

The 6 cubics in (53) span the 4-dimensional vector space of cubics that vanish on p1, …, p6 and can be used to blow up ℙ2 at p1, …, p6 to obtain a cubic surface S. This surface is the closure of

{(a(u),b(u),c(u),d(u),e(u),f(u)):uP2})P5, (54)

and the zero set of the Cremona hexahedral equations

z13+z23+z33+z43+z53+z63=0z1+z2+z3+z4+z5+z6=0a¯z1+b¯z2+c¯z3+d¯z4+e¯z5+f¯z6=0 (55)

where the scalars ā, …, are as in (37). Here, z1, …, z6 are the coordinates of ℙ5.

The Cremona hexahedral form of S can be used to find the 27 lines on S explicitly via expressions for the 45 tritangent planes of S, see [6, Section 4]. In particular the 15 lines that are not part of the Schläfli double six corresponding to the blow up are obtained by permuting the equations

z1+z4=z2+z5=z3+z6=0. (56)

Under the blow up morphism π : S → ℙ2, these lines correspond to specific lines pipj; the one above corresponds to the line between p1 and p2.

We can characterize the rank deficiency of Z6 via the Cremona hexahedral form.

Theorem 6.22

Let x1, …, x6 Px2 and y1, …, y6 Py2 be in general position and let Sx (respectively Sy) be the blow up of Px2 (respectively Py2 ) at x1, …, x6 (respectively y1, …, y6) in Cremona hexahedral form. Then Sx = Sy if and only if either there exists homography H such that Hxi = yi for i = 1, …, 6 or the matrix Z6 is rank deficient.

Proof

If Sx = Sy, then they both contain the 15 lines in equation (56), none of which correspond to the proper transforms of either set of 6 points under their respective blow up morphisms. It follows that either the exceptional lines of x1, …, x6 are exactly the exceptional lines of y1, …, y6 or that the two sets are entirely distinct. Moreover, since each of the 15 lines in (56) have specific intersection conditions, either xi = yi for all i, or {x1, …, x6} and {y1, …, y6} form a Schläfli double six.

In the first case, the exceptional lines of x1, …, x6 are exactly the exceptional lines of y1, …, y6. Then we can define an invertible projective transformation πy πx1 : Px2 Py2 , where πx and πy again refer to the blow up morphisms, such that xiyi for all i. This means that for each i = 1, …, 6, the points xi and yi are related by a homography H.

In the second case the exceptional lines of x1, …, x6 are distinct from the exceptional lines of y1, …, y6. In this case they form a Schläfli double six and by Theorem 6.10, Z6 is rank deficient.

To prove the reverse direction, suppose that Hxi = yi for some homography H. Then Sx = Sy because (āx, … x) ∼ (āy, …, y) by the formula in (37) which means that Sx and Sy have the same Cremona hexahedral form. Similarly, if Z6 is rank deficient then by Theorem 6.10, Px2 blown up at {xi} agrees with Py2 blown up at {yi} and hence Sx = Sy.□

Theorem 6.22 provides a geometric proof of the algebraic characterization of rank deficiency in Theorem 6.7. As before, we index %subscript the cubics in (53) with x or y depending on whether they come from {xi} or {yi}. The Cremona hexahedral equations (55) imply that the two cubic surfaces Sx and Sy are equal if and only if the following polynomials in u (coordinates on Px2 ) and v (coordinates on Py2 ) are identically zero:

a¯xay(v)+b¯xby(v)+c¯xcy(v)+d¯xdy(v)+e¯xey(v)+f¯xfy(v)0a¯yax(u)+b¯ybx(u)+c¯ycx(u)+d¯ydx(u)+e¯yex(u)+f¯yfx(u)0. (57)

In other words,

a¯xz1+b¯xz2+c¯xz3+d¯xz4+e¯xz5+f¯xz6=0for all zSya¯yz1+b¯yz2+c¯yz3+d¯yz4+e¯yz5+f¯yz6=0for all zSx (58)

or equivalently, the matrix

111111a¯xb¯xc¯xd¯xe¯xf¯xa¯yb¯yc¯yd¯ye¯yf¯y (59)

is rank deficient. One can check using Macaulay2 that the ideal of 3 × 3 minors of (59) coincides with the ideal 𝓘nvariants(Z6) of 2 × 2 minors of (38). Therefore, we conclude that Sx and Sy coincide if and only if (39) holds, which is the same as

a¯xb¯xc¯xd¯xe¯xf¯x=λa¯yb¯yc¯yd¯ye¯yf¯y

for some non-zero scalar λ.

This concludes the discussion about 6 points {(xi,yi)}i=16 in general position.

6.2.4 The Proof of Theorem 6.8

We can now use the tools developed in the previous subsection to prove Theorem 6.8 which is the last statement left to prove. This is the case of one ℙ2 having all input points on a line; suppose this is Py2 . We first show that under this assumption, (57) can be simplified using Lemma 3.1. Note that if y1, …, y6 are on a line then for all points v

(ay,by,cy,dy,ey,fy)(v)=(0,0,0,0,0,0), (60)

and for all points v Py2 , using Lemma 3.1,

(ay,by,cy,dy,ey,fy)(v)(Ay,By,Cy,Dy,Ey,Fy) (61)

where Ay, …, Fy are the Joubert invariants introduced in (41). The Joubert invariants are known to satisfy

A+B+C+D+E+F=0A3+B3+C3+D3+E3+F3=0 (62)

see [13, Theorem 9.4.10], and any 5 of them give a complete basis for the space of invariants for 6 ordered points in ℙ1. In particular, if p1, …, p6 ∈ ℙ1 and q1, …, q6 ∈ ℙ1 are two sets of 6 distinct points, then there exists a homography H sending piqi if and only if

(Ap,,Fp)(Aq,,Fq). (63)

We further note that if {yi}i=16 are on a line then āy = … = y = 0 by looking at the definition of [(ij)(kl)(rs)]. We can then simplify (57) to the single equation (42), which was

a¯xAy+b¯xBy+c¯xCy+d¯xDy+e¯xEy+f¯xFy=0.

The following lemma proves the first part of Theorem 6.8.

Lemma 6.23

Let {(xi,yi)}i=16 ⊂ ℙ2 × ℙ2 be such that the points in one2 are on a line. Then Z6 is rank deficient if and only if (42) holds.

Proof

The proof is relatively straightforward. Suppose that y1, …, y6 are on a line. Applying homographies to each side we may assume that the xi’s are all finite and that yi = (yi1, 1, 0). Then

Z6=x11y11x110x12y11x120y1110x21y21x210x22y21x220y2110x31y31x310x32y31x320y3110x41y41x410x42y41x420y4110x51y51x510x52y51x520y5110x61y61x610x62y61x620y6110 (64)

and Z6 is rank deficient if and only if its only obviously non-zero 6 × 6 minor det(M) = 0. Furthermore, a quick computation will give

a¯xAy+b¯xBy+c¯xCy+d¯xDy+e¯xEy+f¯xFy=24det(M) (65)

yielding the desired result.□

Before we prove the second part of Theorem 6.8, we compute an example to illustrate the statement.

Example 6

Consider the following points in which the x points are in general position and the y points are all distinct and contained in a line. This can be seen as a configuration in ℙ2 × ℙ1 so we fix

x1=(0,0,1)y1=(0,1)x4=(1,1,1)y4=(4,1)x2=(1,0,1)y2=(1,1)x5=(3,5,1)y5=(8,1)x3=(0,1,1)y3=(3,1)

and leave x6, y6 as symbolic finite points. This creates a 6 × 6 matrix

Z6=00000111001100313141414124340581x61y61x61x62y61x62y611. (66)

(The reader may also view the yi as points on the line at infinity in ℙ2 where the third coordinate is 0; the non-zero part of the corresponding 6 × 9 matrix Z6 is (66).)

The matrix Z6 in (66) is rank deficient exactly when its determinant, a quadratic polynomial in the coordinates of x6 and y6, is 0. If we arbitrarily fix x6 = (2, 11, 1), then det(Z6) = –918y61 + 2942, and y6 = (2942, 918) is the unique point y6 such that the configuration is rank deficient.

If we consider only the first 5 point pairs we fixed initially, there exists a unique projective transformation

T=146294013510911 (67)

such that Txiyi for all i = 1, …, 5. Moreover, we can check that Tx6 = (–2942, –918) ∼ y6. This transformation T has center u = (–1617, –803, 11888). This point actually has some further significance, because we can check that (ax(u), …, fx(u)) ∼ (48079, –55599, –88559, –17265, 22529, 90815) ∼ (Ay, …, Fy).

The following now finishes the proof of Theorem 6.8.

Proof of Theorem 6.8

(second part). Let S be the blow up of Px2 in x1, …, x6 in Cremona hexahedral form, and recall (55) and (62). It follows that (42) holds if and only if (Ay, …, Fy) ∈ S, which holds if and only if there exists u Px2 such that

(ax(u),,fx(u))(Ay,,Fy). (68)

We first show that if Z6 is rank deficient then a projection exists. Let Px2 be a line such that u. We then construct a projection T′ : Px2 with center u as in the proof of Theorem 5.2 and Figure 2. This projection has T′(xi) = ai for all i, where the ai are defined as in Figure 2. Then

(Aa,,Fa)(ax(u),,fx(u))(Ay,,Fy). (69)

Since the Joubert invariants generate the space of all invariants for 6 ordered points on a line, it follows that there exists a homography H : Py1 such that Hai = yi. We conclude that T = HT′ is the desired projection T : Px2 Py1 .

To prove the reverse direction, suppose that such a projection T exists. We can use homography on each side to fix u = (0, 0, 1) and to fix the image line at infinity. Equivalently, pick homographies on each side such that

T=100010. (70)

It then follows that xi = (yi1, yi2, bi) for all i, where the bi are unknown scalars. The matrix Z6 then becomes

y112y11y12y11y12y122y11b1y12b1y212y21y22y21y22y222y21b2y22b2y312y31y32y31y32y322y31b3y32b3y412y41y42y41y42y422y41b4y42b4y512y51y52y51y52y522y51b5y52b5y612y61y62y61y62y622y61b6y62b6 (71)

which is clearly rank deficient. This concludes the proof of Theorem 6.8.□

Funding statement: A.E. is supported by NSF CCF 2110075.

Acknowledgements

We thank Jarod Alper, Timothy Duff, and Giorgio Ottaviani for helpful discussions.

  1. Communicated by: D. Plaumann

References

[1] M. Bertolini, L. Magri, C. Turrini, Critical loci for two views reconstruction as quadratic transformations between images. J. Math. Imaging Vision 61 (2019), 1322–1328. MR4017694 Zbl 1482.68246Search in Google Scholar

[2] M. Bråtelund, Critical configurations for three projective views. Math. Scand. 129 (2023), 401–444. MR4667876 Zbl 07849503Search in Google Scholar

[3] M. Bråtelund, Critical configurations for two projective views, a new approach. J. Symbolic Comput. 120 (2024), Paper No. 102226, 22 pages. MR4583113 Zbl 07725351Search in Google Scholar

[4] A. Buckley, T. Košir, Determinantal representations of smooth cubic surfaces. Geom. Dedicata 125 (2007), 115–140. MR2322544 Zbl 1117.14038Search in Google Scholar

[5] P. Bürgisser, F. Cucker, Condition. The geometry of numerical algorithms, volume 349 of Grundlehren Math. Wiss. Springer 2013. MR3098452 Zbl 1280.65041Search in Google Scholar

[6] A. B. Coble, Point sets and allied Cremona groups. I. Trans. Amer. Math. Soc. 16 (1915), 155–198. MR1501008 Zbl 02618070Search in Google Scholar

[7] E. Connelly, R. Thomas, C. Vinzant, The geometry of rank drop in a class of face-splitting matrix products: Part II. Adv. Geom. 24 (2024), 395–420.Search in Google Scholar

[8] J. W. Demmel, The geometry of ill-conditioning. J. Complexity 3 (1987), 201–229. MR907197 Zbl 0641.65042Search in Google Scholar

[9] J. W. Demmel, On condition numbers and the distance to the nearest ill-posed problem. Numer. Math. 51 (1987), 251–289. MR895087 Zbl 0597.65036Search in Google Scholar

[10] J. W. Demmel, Applied numerical linear algebra. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA 1997. MR1463942 Zbl 0879.65017Search in Google Scholar

[11] H. Derksen, G. Kemper, Computational invariant theory, volume 130 of Encyclopaedia of Mathematical Sciences. Springer 2015. MR3445218 Zbl 1332.13001Search in Google Scholar

[12] I. Dolgachev, D. Ortland, Point sets in projective spaces and theta functions. Astérisque no. 165 (1988), 210 pages (1989). MR1007155 Zbl 0685.14029Search in Google Scholar

[13] I. V. Dolgachev, Classical algebraic geometry. Cambridge Univ. Press 2012. MR2964027 Zbl 1252.14001Search in Google Scholar

[14] D. Eisenbud, Commutative algebra. Springer 1995. MR1322960 Zbl 0819.13001Search in Google Scholar

[15] H. Fan, J. Kileel, B. Kimia, On the instability of relative pose estimation and RANSAC’s role. In: Proc. IEEE/CVF Conference on Computer Vision and Pattern Recognition, 2022, 8925–8933.Search in Google Scholar

[16] D. R. Grayson, M. E. Stillman, Macaulay2, a software system for research in algebraic geometry. www.math.uiuc.edu/Macaulay2/.Search in Google Scholar

[17] R. Hartley, F. Kahl, Critical configurations for projective reconstruction from multiple views. International Journal of Computer Vision 71 (2007), 5–47. Zbl 1477.68366Search in Google Scholar

[18] R. Hartley, A. Zisserman, Multiple view geometry in computer vision. Cambridge Univ. Press 2003. MR2059248 Zbl 1072.68104Search in Google Scholar

[19] R. Hartshorne, Algebraic geometry. Springer 1977. MR463157 Zbl 0367.14001Search in Google Scholar

[20] D. Hilbert, S. Cohn-Vossen, Geometry and the imagination. Chelsea Publishing Co., New York 1952. MR46650 Zbl 0047.38806Search in Google Scholar

[21] F. Kahl, R. Hartley, Critical curves and surfaces for euclidean reconstruction. In: Computer Vision 2002: 7th European Conference on Computer Vision, Copenhagen 2002, Proc. Part II, Lect. Notes Comput. Sci. 2351, Springer 2002, 447–462. Zbl 1039.68658Search in Google Scholar

[22] C. G. Khatri, C. R. Rao, Solutions to some functional equations and their applications to characterization of probability distributions. Sankhyā Ser. A 30 (1968), 167–180. MR238416 Zbl 0176.49004Search in Google Scholar

[23] Q. T. Luong, O. D. Faugeras, The fundamental matrix: Theory, algorithms, and stability analysis. International Journal of Computer Vision 17 (1996), 43–75.Search in Google Scholar

[24] G. Ottaviani, Five lectures on projective invariants. Rend. Semin. Mat. Univ. Politec. Torino 71 (2013), 119–194. MR3345059 Zbl 1312.13012Search in Google Scholar

[25] A. Pryhuber, R. Sinn, R. R. Thomas, Existence of two view chiral reconstructions. SIAM J. Appl. Algebra Geom. 6 (2022), 41–76. MR4387191Search in Google Scholar

[26] J. G. Semple, G. T. Kneebone, Algebraic projective geometry. Oxford Univ. Press 1998. MR1651410 Zbl 0932.51008Search in Google Scholar

[27] V. I. Slyusar, A family of face products of matrices and its properties (Russian). Kibernet. Sistem. Anal. no. 3 (1999), 43–49, 189. English translation in Cybernet. Systems Anal. 35 (1999), no. 3, 379–384. MR1728929 Zbl 0984.15015Search in Google Scholar

[28] R. Sturm, Die Lehre von den geometrischen Verwandtschaften. Dritter Band: Die eindeutigen linearen Verwandtschaften zwischen Gebilden dritter Stufe. Teubner, Leipzig 1909. Zbl 39.0604.01Search in Google Scholar

[29] B. Sturmfels, Algorithms in invariant theory. Springer 2008. MR2667486 Zbl 1154.13003Search in Google Scholar

[30] C. F. Van Loan, The ubiquitous Kronecker product. J. Comput. Appl. Math. 123 (2000), 85–100. MR1798520 Zbl 0966.65039Search in Google Scholar

[31] T. Werner, Constraint on five points in two images. In: Proc. IEEE Conference on Computer Vision and Pattern Recognition 2 (2003), 203–208.Search in Google Scholar

A Appendix: Facts about cubic surfaces

This is a collection of classical facts about cubic surfaces needed in Subsection 6.2. See [19, V. Chapter 4] for a modern reference and [6] or [20] for classical references.

Lemma A.1

Let 𝓢 ⊂ ℙ3 be a smooth cubic surface. Then the following hold.

  1. 𝓢 can be obtained by blowing up 6 points p1, …, p6 ∈ ℙ2 in general position. The exceptional curves of p1, …, p6 in the blow up morphism π : 𝓢 → ℙ2 are 6 lines 1, …, 6 on 𝓢.

  2. The cubic surface 𝓢 has exactly 27 (complex) lines. They are

    1. the exceptional lines 1, …, 6,

    2. the strict transforms ij of the lines in2 passing through pi, pj for 1 ≤ i < j ≤ 6 (15 of them), and

    3. the strict transforms j of the conics in2 passing through five of the pi for ij, j = 1, …, 6 (6 of them).

  3. The two sets of lines {1, …, 6} and {1,,6} form a Schäfli double six on 𝓢. There are 36 Schläfli double sixes on 𝓢. Each double six is uniquely determined by three of the line pairs (i, i ).

  4. Any set of 6 mutually skew lines among the 27 lines on 𝓢 can play the role of 1, …, 6. Precisely, if {g1, …, g6} is a set of 6 mutually skew lines among the 27 lines, then there is another morphism π′ : 𝓢 → ℙ2, making 𝓢 isomorphic to2 with six points q1, …, q6 in general position blown up such that g1, …, g6 are the exceptional curves of q1, …, q6 under π′.

We now note several facts about double sixes in ℙ3 (or more generally line configurations in ℙ3).

Lemma A.2

  1. Each double six in3 is contained in a unique non-singular cubic surface and thus forms part of the set of 27 lines on that surface.

  2. Any set of 6 mutually skew lines among 27 lines that satisfy the incidence relations is one half of a unique Schläfli double six and determines the other lines and the cubic surface.

Next we talk about determinantal representations of a cubic surface.

Definition A.3

A determinantal representation of a cubic surface 𝓢 defined by a cubic equation F(z) = F(z0, z1, z2, z3) = 0 is a 3 × 3 matrix of linear forms

M(z)=M(z0,z1,z2,z3)=z0M0+z1M1+z2M2+z3M3 (72)

such that det(M(z)) = λF(z) for some λ ≠ 0. Two representations M(z) and M′(z) are equivalent if M′(z) = AM(z)B with A, B ∈ GL(3).

Note that if M(z) is a determinantal representation of 𝓢, then so is M(z). We present the following facts about determinantal representations of a cubic surface; see [4].

Lemma A.4

  1. Every smooth cubic surface 𝓢 has a determinantal representation.

  2. The equivalence classes of determinantal representations of 𝓢 are in bijection with sets of 6 mutually skew lines on 𝓢 that are the exceptional curves of the blow up of2 at the 6 points that give 𝓢.

  3. If the determinantal representation M of 𝓢 corresponds to the skew lines 1, …, 6 on 𝓢, then M corresponds to another 6 skew lines 1,,6 on 𝓢 such that the pair of 6 lines form a Schläfli double six on 𝓢. Thus there are 72 equivalence classes of determinantal representations of 𝓢, one for each half of the 36 Schläfli double sixes on 𝓢.

  4. The choice of a Schläfli double six 1, …, 6, 1,,6 on 𝓢 is equivalent to the choice of a pair of equivalence classes of determinantal representations of 𝓢, namely[M] and [M]. (A smooth cubic surface cannot have a symmetric determinantal representation and hence [M] ≠ [M].)

Here are various recipes to explicitly realize some of the above assertions.

  1. If 𝓢 is the blow up of ℙ2 at the 6 points p1, …, p6, then 𝓢 can be constructed as follows. Pick a basis {f0, f1, f2, f3} of the 4-dimensional vector space of cubics through p1, …, p6 and map

    uP2(f0(u):f1(u):f2(u):f3(u))P3. (73)
  2. The variety of 〈f0, …, f3〉 is {p0, …, p6}. By the Hilbert–Burch theorem [14, Section 20.4] there is a 3 × 4 matrix L(u) of linear forms in u0, u1, u2 whose maximal minors are f0, …, f3. The matrix L(u) generically has rank 3 and has rank 2 exactly when u = pi for some i = 1, …, 6. Define a 3 × 3 matrix M(z) of linear forms in z0, z1, z2, z3 by the relation

    M(z)u0u1u2=L(u)z0z1z2z3. (74)

    Then M(z) is a determinantal representation of 𝓢.

  3. The blow up of ℙ2 at p1, …, p6 can be seen via (74). If ū ∈ ℙ2 is not one of p1, …, p6, then some fi(ū) ≠ 0 and rank(L(ū)) = 3. Therefore, the null space of L(ū) is a unique point ∈ ℙ3 and M()ū = L(ū) = 0. This means that det(M()) = 0 and ∈ 𝓢. If is a generic point on 𝓢, then rank(M()) = 2, and the blow up morphism π sends to the unique point ū ∈ ℙ2 that generates the right null space of M().

  4. The 6 exceptional curves of the blow up are precisely

    i:={zP3:M(z)pi=0} for i=1,,6. (75)

    By what is stated above, rank(L(u)) = 2 if and only if u = pi for some i = 1, …, 6 if and only if (via (74)) M(z)u = L(u)z = 0 is a system of 2 linearly independent equations in z (i.e., defines a line in ℙ3). This line is entirely on 𝓢 since det(M(z)) = 0 if M(z)pi = 0.

  5. Corresponding results hold for the determinantal representation M(z) of 𝓢.

    1. To obtain the points q1, …, q6 ∈ ℙ2 whose blow up is 𝓢, compute L′(u) via

      M(z)u=L(u)z. (76)

      The variety of the maximal minors of L′(u) is {q1, …, q6}.

    2. A point ȳ ∈ ℙ2 not equal to any qi is sent to the unique point in ℙ3 in the right nullspace of L′(ȳ). This means L′(ȳ) = M() ȳ = 0 and ∈ 𝓢 since det(M()) = 0.

    3. The morphism π′ : 𝓢 → ℙ2 sends to the unique point ȳ ∈ ℙ2 that generates the right nullspace of M(), or equivalently, the left nullspace of M().

    4. The exceptional curves of the blow up are the lines

      j:={zP3:qjM(z)=0} for j=1,,6.

The following theorem follows from the above facts.

Theorem A.5

Given a determinantal representation M(z) of a smooth cubic surface 𝓢 ⊂ ℙ3, there exists 6 points (pi, qi) ∈ ℙ2 × ℙ2 such that 𝓢 is the blow up of the first2 at {pi}i=16 and also the blow up of the second2 at {qj}j=16. The exceptional curves of the two blow ups form a Schläfli double six on 𝓢:

i:={zP3:M(z)pi=0}fori=1,,6,j:={zP3:qjM(z)=0}forj=1,,6.
Received: 2023-09-12
Revised: 2024-03-06
Published Online: 2024-08-05
Published in Print: 2024-07-26

© 2024 Walter de Gruyter GmbH, Berlin/Boston, Germany

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 25.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/advgeom-2024-0016/html
Scroll to top button