You are on page 1of 15

arXiv:1505.04204v1 [math.

AG] 15 May 2015

TRUNCATED MODULES AND


LINEAR PRESENTATIONS OF VECTOR BUNDLES
ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA
Abstract. We give a new method to construct linear spaces of matrices of constant rank, based
on truncated graded cohomology modules of certain vector bundles. Our method allows one to
produce several new examples, and provides an alternative point of view on the existing ones.

Introduction
A space of matrices of constant rank is a vector subspace V , say of dimension n + 1, of the
set Ma,b (k) of matrices of size a b over a field k, such that any element of V \ {0} has fixed
rank r. It is a classical problem, rooted in work of Kronecker and Weierstrass, to look for such
spaces of matrices, and to give relations among the possible values of the parameters a, b, r, n. The
main difficulty here is to construct matrices with small constant corank with respect to a and b,
especially so when n grows, since n > 3 is already hard. One can see V as an a b matrix whose
entries are linear forms (a linear matrix), and interpret the cokernel as a vector space varying
smoothly (i.e. a vector bundle) over Pn . In [EH88, IL99, Syl86] the relation between matrices of
constant rank and the study of vector bundles on Pn and their invariants was first studied in detail.
This interplay was pushed one step further in [BFM13, BM14], where the matrix of constant rank
was interpreted as a 2-extension from the two vector bundles given by its cokernel and its kernel.
This allowed to construct skew-symmetric matrices of linear forms in 4 variables of size 14 14
and corank 2, beyond the previous record of [Wes96].
In this paper we turn the tide once again and introduce a new effective method to construct linear
matrices of constant rank. This should be seen as a parallel of [BFM13] which complements and
explains the techniques used there, relying on commutative algebra rather than derived categories.
The starting point of our analysis is that linear matrices of relatively small size can be cooked
up with two ingredients, namely two finitely generated graded modules E and M over the ring
R = k[x0 , . . . , xn ], admitting a linear resolution up to a certain step. Here, the module M should
be thought of as a small modification of E, namely M should be Artinian. Then, under suitable
conditions on the Betti numbers of E and M , a surjective map : E M induces a resolution
of ker() which is also linear, and of smaller size, as the presentation matrix of M is subtracted
from that of E. The key idea here is that, in order for E to fit our purpose, it is necessary to
truncate E above a certain range, typically the regularity of E, which ensures linearity of the
resolution, while leaving the rank of the matrix presenting E unchanged. All this is stated in a
more precise form in Theorem 1.1 and in the auxiliary results of 1.
To connect this result with linear matrices of constant rank, one takes E to be the module of
global sections of a vector bundle E over Pn , which guarantees that the presentation matrix of E,
2010 Mathematics Subject Classification. 13D02, 16W50, 15A30, 14J60.
Key words and phrases. Graded truncated module, linear presentation, matrix of constant rank, vector bundle,
instanton bundle.
Research partially supported by MIUR funds, PRIN 2010-2011 project Geometria delle variet`
a algebriche,
FIRB 2012 Moduli spaces and Applications, GEOLMI ANR-11-BS03-0011, J. Draismas Vidi grant from the
Netherlands Organization for Scientific Research (NWO) and Universit`
a degli Studi di Trieste FRA 2011 project
Geometria e topologia delle variet`
a. All authors are members of GNSAGA.
1

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

as well as that of any higher truncation of E, actually has constant rank. This is the content of
Theorem 2.1. Then we analyze basic Artinian modules M (especially those of length 1 and 2) in
order to obtain interesting matrices via this method. Note that the choice of where one needs to
truncate is arbitrary here; in this sense we define a tree associated with E, rooted in E, whose
edges are the allowed levels where we may truncate E, and whose leaves are the possible forms of
matrices whose cokernel is E.
One of the goals of this paper is to provide a list of matrices of constant rank arising from vector
bundles over projective spaces. We concentrate here on the most classical constructions (instantons,
null-correlation bundles and so forth). To our knowledge however, among the examples of matrices
of constant rank constructed from these bundles, very few were known before. Also, the rank and
size of the matrices appearing in many of these examples are in a range that does not seem easy
to reach with previously known techniques. All this is developed in 2.
In the last section we examine the conditions needed for a constant rank matrix A constructed with our method to be skew-symmetric in a suitable basis (we call such matrix skewsymmetrizable). This is tightly related with the results of [BFM13]: there, some skew-symmetric
linear matrices of constant rank were constructed starting with a vector bundle together with a
piece of homological data, namely a certain extension class involving the cokernel bundle of A.
This time our approach relies on truncated modules rather than derived categories. Our new
technique has the advantage of being algorithmic, and can be actually implemented in a very
efficient way, as computations sometimes work even beyond their theoretical scope. This not only
provides a detailed explanation of the algorithm appearing in [BFM13, Appendix A], but allows
to construct infinitely many examples of skew-symmetric 10 10 matrices of constant rank 8 in 4
variables; up to now, the only example of such being those of [Wes96].
Finally, let us note that all these examples, and many more, can be explicitly constructed thanks
to the Macaulay2 package ConstantRankMatrices implementing our algorithms. The interested
reader can find it at the website www.paololella.it/EN/Publications.html.
1. Main results
1.1. Notation and preliminaries. Let R := k[x0 , . . . , xn ] be a homogeneous polynomial ring
over an algebraically closed field k of characteristic other than 2. The ring R comes with a grading
R = R0 R1 . . ., with R0 = k, and is generated by R1 as a k-algebra.
All R-modules here are finitely generated and graded. If M is such a module, and p, q are integer
numbers, we denote by M p the p-th graded component of M , so that M = p M p , and by M (q)
the qth shift of m, defined by the formula M (q)p = M p+q . Finally, the module M >m = p>m M p
is the truncation of M at degree m.
A module M is free if M i R(qi ) for suitable qi . Given any other finitely generated graded
R-module N , we write Hom(M , N ) for the set of homogeneous maps of all degrees, which is again
a graded module, graded by the degrees of the maps. Since graded free resolutions exist, this
construction extends to a grading on the modules ExtqR (M , N ). The same holds true for M N
and Tor(M , N ). For any R-module M , we set i,j (M ) := dimk Tori (M , k)j .
For p > 0, the Hilbert function dimk M p is a polynomial in p. The degree of this polynomial,
increased by 1, is dim M , the dimension of M . The degree of M is by definition (dim M !) times
the leading coefficient of this polynomial.
We say that M has m-linear resolution over R if the minimal graded free resolution of M reads:
R(m 2)2 R(m 1)1 R(m)0 M 0
for suitable integers i . In other words, M has a m-linear resolution if M r = 0 for r < m, M
is generated by M m , and M has a resolution where all the maps are represented by matrices of
linear forms. If only the first k maps are matrices of linear forms then M is said by m-linear
presented up to order k, or just linearly presented when k = 1.

A module M is m-regular if, for all p, the local cohomology groups Hpm (M )r vanish for r =
m p + 1 and also H0m (M )r = 0 for all r > m + 1. The regularity of a module is denoted by
reg(M ). Note that M >m always has m-linear resolution if m is large enough.
1.2. Main Theorem. Here comes our main result on the linear presentation of modules.
Theorem 1.1 (The Brunch Theorem). Let E be an R-module, and let m > reg(E) be an integer. Let M be an Artinian R-module, m-linearly presented up to order 2, and assume that
i,m+i (E >m ) > i,m+1 (M ) for i = 0, 1, 2. Assume moreover that there is a surjective morphism
: E >m M . Then, the kernel F := Ker is m-linearly presented.
Proof. Since is surjective, we have a short exact sequence of graded modules:

M 0
0 F E >m
This induces the long exact sequence (see [Pee11, Theorem 38.3]):

(1.1)

Tor2 (E >m , k)

Tor2 (M , k)

Tor1 (F , k)

Tor1 (E >m , k)

Tor1 (M , k)

Tor0 (F , k)

Tor0 (E >m , k)

Tor0 (M , k)

where the modules Tori (, k) are graded of finite length and the dimensions of the homogeneous
pieces are equal to the Betti numbers of the minimal free resolutions of the modules [Pee11,
Theorem 11.2].
By assumption, the module E >m has regularity m. By [EG84, Theorem 1.2], M >m has mlinear resolution if and only if M is m-regular. So, there exist integers 0 , 1 , . . . such that E >m
has the m-linear resolution:
R(m 2)2 R(m 1)1 R(m)0 E >m 0,
so that i,m+i (E >m ) = i and i,j (E >m ) = 0 for all j 6= m + i.
The first steps of the minimal graded free resolution of M are:
R(m 2)2 R(m 1)1 R(m)0 M 0.
This tells that, for any i = 0, 1, 2:
i,m+i (M ) = i ,

i,j (M ) = 0, if j > m + i.

The same vanishing holds for all i if j < m + i.


Let us now consider the long exact sequences induced by (1.1) on each degree. Obviously, for
j < m, i,j (E >m ) = i,j (M ) = 0 implies i,j (F ) = 0. In the case j = m, one has:
0 Tor0 (M , k)m Tor0 (E >m , k)m Tor0 (F , k)m 0
i.e. 0,m (F ) = 0,m (E >m ) 0,m (M ) = 0 0 . The degree m + 1 part of (1.1) is:
0 Tor1 (F , k)m+1 Tor1 (E >m , k)m+1 Tor1 (M , k)m+1 Tor0 (F , k)m+1 0
and thus:
1,m+1 (F ) 0,m+1 (F ) = 1,m+1 (E >m ) 1,m+1 (M ).
By hypothesis, 1,m+1 (E >m ) 1,m+1 (M ) = 1 1 > 0 and by the minimality of the resolution
of F we have 0,m+1 (F ) = 0 and 1,m+1 (F ) = 1 1 . For degree j = m + 2, we have:
0 Tor2 (F , k)m+2 Tor2 (E >m , k)m+2 Tor2 (M , k)m+2 Tor1 (F , k)m+2 0
and we repeat the same argument as above to conclude 1,m+2 (F ) = 0 and 2,m+2 (F ) = 2 2 .
Finally, for j > m + 2, 1,j (F ) = 0 as 1,j (E >m ) = 2,j (M ) = 0. Hence, the presentation of F

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

is linear as the modules Tor0 (F , k) and Tor1 (F , k) have length one with Tor0 (F , k)m 6= 0 and
Tor1 (F , k)m+1 6= 0.

Remark 1.1. The name Brunch Theorem for Theorem 1.1 comes from the fact that a rough
draft of its proof came out exactly during a Saturday morning brunch in Turin, in June 2014.
1.3. A more detailed analysis. Keeping in mind our goal of constructing explicit examples of
constant rank matrices, we now want to investigate the features of modules satisfying Theorem 1.1s
hypotheses. We take into consideration Artinian modules of length 1 and 2, and seek numerical
ranges for Betti numbers satisfying the hypotheses of Theorem 1.1. Note that this does not
guarantee the existence of the surjective morphism .
Let us suppose that the module M is generated in degree m and has length + 1. Its Hilbert
series is tm (h0 + h1 t + . . . + h t ), and it can be expressed as a function of its Betti numbers:

n+1
P
P 
(1)i j ij tj
.
tm (h0 + h1 t + . . . + h t ) = i=0
(1 t)n+1
In our setting one has:
(1 t)n+1 tm (h0 + h1 t + . . . + h t ) = 0 tm 1 tm+1 + 2 tm+2 + . . .
and from this we deduce:
(
0 = h 0
1 = (n + 1)h0 h1

h 0 = 0
h1 = (n + 1)0 1

(n + 1)0 1 > 0.

From explicit computations one sees that:


Proposition 1.2. Let M be a finitely generated graded module of length 2 over R = k[x0 , . . . , xn ],
satisfying all hypotheses of Theorem 1.1. Then the last piece of its resolution has the form:
R(r 2)(n+1)1 (

)0 R(r 1)1 R(r)0 M 0,

n+2
2

with

2n+4
3 0

6 1 6 (n + 1)0 .

Remark 1.2. The resolution of Artinian modules of length > 3 is much more involved. In particular,
for length > 3, the Hilbert function and the first two Betti numbers do not determine the shape
of the free module of homological position 2.
In what follows we will need the next result:
Lemma 1.3. Let M be a finitely generated graded R-module and let m > reg(M ) be an integer
number. The truncated module M >m is m-regular, and assume that it has linear resolution:
0 R(m n)n,m R(m i)i,m R(m)0,m M >m 0.
Then the truncated module M >m+k , k > 1, has regularity m + k and linear resolution:
0 R(m k n)n,m+k R(m k)0,m+k M >m+k 0
with
(i)

(i)

(i)

i,m+k = ak 0,m ak1 1,m + . . . + (1)n akn n,m =

n
X

(i)

(1)j akj j,m ,

j=0

(i) 
ak

where for all i = 0, . . . , n, the sequence


belongs to the set of recursive sequences:




n


X
n+1

akj .
(1)j
RSn := (ak ) ak+1 =


j+1
j=0

5
(i) 

More in detail, ak
(1.2)

(i)

a1 =

is defined by the initial values:



n+1
,
i+1

(i)

ai = (1)i

(i)

aj = 0 for j 6= 1, i and j < n.

and

Proof. Looking at the exact sequence 0 M >m+k+1 M >m+k M m+k 0, one works
repeatedly with the long exact sequence (1.1).

By induction on n one can also prove:
Lemma 1.4. Any recursive sequence (ak ) RSn has a degree n polynomial as its generating func(i) 
tion. In particular, the generating function of the sequence ak defined in (1.2) is the following:
pn(i) (k)

 

n k+n1 k+n
=
.
i
n
k+i
(0)

Remark that in the case i = 0 one gets that pn (k) =

k+n
n


.

Example 1.1. Consider the Koszul complex of R = k[x0 , x1 , x2 , x3 ] as resolution of the irrelevant
ideal m = (x0 , x1 , x2 , x3 ). It reads:
0 R(4) R(3)4 R(2)6 R(1)4 m 0.
By Lemma 1.3 and 1.4, the resolution of m>1+k for every k > 0 is
0 R(4 k)3 (k) R(3 k)2 (k) R(2 k)1 (k) R(1 k)0 (k) m>1+k 0,
where
1 3
k +
6
1
2 (k) = k 3 +
2

0 (k) =

3 2 13
k + k + 4,
2
3
7 2
k + 7k + 4,
2

1 3
19
k + 4k 2 + k + 6,
2
2
1 3
11
2
3 (k) = k + k + k + 1.
6
6
1 (k) =

Therefore, we can predict the resolution of m>10 :


0 R(13)220 R(12)715 R(11)780 R(10)286 m>10 0.

2. Linear presentation of vector bundles


We now focus our attention on the case where the graded R-modules are cohomology modules
of vector bundles over the projective space Pn = Proj R.
Theorem 2.1. In the setting of Theorem 1.1, suppose furthermore that E = H0 (E) is the graded
module of sections of a rank r vector bundle E on Pn . Set a = 0,m (E >m ) 0,m (M ) and
b = 1,m+1 (E >m ) 1,m+1 (M ). Then the presentation matrix A of F is linear of size a b and
e of F is isomorphic to E.
corank r. Moreover the sheafification F := F
Proof. The fact that the presentation matrix of F is linear is consequence of Theorem 1.1.

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

For i > 0, set i = i,m+i (E >m ) and i = i,m+i (M ). Then, from the proof of Theorem 1.1,
we extract the next commutative exact diagram.
0
A

R(m)0 0

R(m 1)1 1

R(m 2)2

R(m 1)1

R(m)0

E >m

R(m 2)2

R(m 1)1

R(m)0

Therefore, the matrix A presenting F is linear of size a b, where a = 1 1 and b = 0 0 .


Note that this holds independently of the fact that E is locally free.
The fact that A has constant corank r follows from the observation that the modules E and F
differ by the Artinian module M only, hence their induced coherent sheaves on Pn are isomorphic.
e is a locally free sheaf of rank r, i.e. the restriction of the presentation matrix
In particular, F = F
A of F has constant corank r.


Let us spell out clearly our strategy to construct linear matrices of constant rank based on this
result. Suppose that, for a given 3-uple of integers (, a, b) with < a 6 b, we want to construct
an a b matrix of linear forms of constant rank in the polynomial ring with n + 1 variables.
Then we have to search for an a b matrix of linear forms presenting a vector bundle E of rank
r = a on Pn . In general, if E is a vector bundle of rank r, its module E of global sections
will not be linearly presented. But we are free to truncate E in such a way that it is m-linearly
presented. By Theorem 2.1, the presentation matrix will indeed have corank r. However, it is
unlikely that its size equals a b. Here is where Theorem 1.1 and 2.1 come into play, as we may
remove a linearly presented Artinian module from our truncation of E in order to reduce the size
of our presentation matrix, and hopefully arrive at size a b.
Note that, once given E, there are infinitely many truncations of E = H0 (E), and for each
truncation E >k one can look at the finitely many Artinian modules of length at most 2 satisfying
the assumption of Theorem 1.1. We illustrate these possibilities with a 2-level infinite tree structure.
The root of the tree is the vector bundle E, the elements at level 1 are all possible truncations
E >reg E+k , k N, and the leaves are possible sizes of linear matrices of constant rank produced
by the algorithm. (Explicit examples can be found in Figures 1-5.) We emphasize that in order to
determine such a tree, one only needs to compute explicitly the first truncation E >reg E with linear
resolution, and then apply Lemma 1.3 and Lemma 1.4. The existence of the matrix is guaranteed
only in cases where there are surjective morphisms from E >reg E to the Artinian module.

Example 2.1 (Line bundles). To give a first application of Theorem 2.1, let E be a line bundle
OPn (l) over Pn . Then, Figure 1 shows how to obtain from Theorem 2.1 matrices of constant
rank min{a, b} 1. Clearly these matrices can constitute building blocks to construct bigger size
matrices, such as square matrices of size 10, 12 and 14 of rank 8, 10 and 12, and also skew-symmetric
matrices of size a + b with rank 2a 2 (assuming a 6 b).
2.1. Steiner bundles, linear resolutions and generalizations. In classical literature a vector bundle E on Pn having a linear resolution of the form:
0 OPn (1)s OPs+r
E 0,
n
with s > 1 and r > n integers, is called a (rank r) Steiner bundle. This motivates the following:

E = OP2 (l)

E l

k=0

0
1

E l+1

k=1

l + 1

0
3

1
3

2
1

E l+2

k=2

l + 2

0
6

1
8

l + 3

0
10

1
15

1
3

2
3

3
1

5 5 matrix of rank 4

2
3

l + 3

0
1

1
3

2
3

3
1

l + 3

0
2

1
6

2
6

3
2

l + 3
l + 4

0
3

1
8

2
6

E l+3

k=3

0
1

l + 2

9 12 matrix of rank 8

2
6

8 9 matrix of rank 7

7 7 matrix of rank 6

E l+4

k=4
l + 4

0
15

1
24

0
1

l + 4

1
3

2
3

3
1

14 21 matrix of rank 13

2
10
l + 4

0
2

1
6

2
6

3
2

l + 4
l + 5

0
3

1
8

2
6

l + 4

0
3

1
9

2
9

3
3

1
11

2
9

13 18 matrix of rank 12

12 16 matrix of rank 11
12 15 matrix of rank 11
0
4

l + 4
l + 5

3
1
1

11 13 matrix of rank 10

Figure 1. A piece of the tree associated with a line bundle over P2 . Each
truncation is described by the Betti table of its resolution, and the arrows from
the truncated modules to possible sizes of constant rank matrices are labeled with
the Betti tables of the Artinian modules satisfying the assumptions of Theorem
1.1 and Proposition 1.2. In all cases pictured above one can show the existence of
the surjective morphism, and thus of the matrix.
(m)

Definition 2.1. Let r > n and s > 1 be integer numbers. The cokernel Es,r of a generic
morphism:

OPs+r
OPn (m 1)s
n
is a vector bundle on Pn , that we call generalized Steiner bundle.
(0)

Classical Steiner bundles are of the form Es,r . Given a generalized Steiner bundle, the graded
(m)
module E (m) := H0 (Es,r ) has the following resolution:
(2.1)

0 R(m 1)s Rs+r E (m) 0,

and its regularity is exactly equal to m. Remark though that the module E (m) does not fit the
bill for our purposes: first of all for m 6= 0 the resolution is not linear, and even in the case m = 0

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

one has Tor2 (E (0) , k)2 = 0. For this reason we need to truncate it in higher degree, as explained
in the following:
(m)

Lemma 2.2. Let Es,r be a rank r generalized Steiner bundles, with graded modules of sections
(m)
(m)
E (m) = H0 (Es,r ). The linear resolutions of the truncation E >m is of the form:
(m)

(m)

(m)

0 R(m i)i,m R(m)0,m E >m 0


where:
(m)

(i)
i,m = p(i)
n (m)(s + r) pn (1)s.

(2.2)

(m)

Proof. Looking at the homogeneous piece of degree m + i of the resolution of E >m


(m)

(m) 
m+i

(m)

i,m
0,m
0 R(m i)m+i
R(m)m+i
E >m

we deduce that:
(m)

i,m =

i
X

(1)j1

n+j
n

j=1

(m) 
m+i

Since E >m

degree m + i of

= E (m)

(m)
E >m

m+i

0,

 (m)
(m) 
ij,m + (1)i dimk E >m m+i .

, i > 0, we compute the dimension of the homogeneous piece of

from the simpler resolution (2.1):


(m) 
m+i

dimk E >m

= dimk E (m)

m+i


(s + r)

n+m+i
n

n+i1
n


s.

We now proceed by induction on i; for i = 0, we have:




(m) 
(m)
(0)
(0)
0,m = dimk E >m m = n+m
(s + r) n1
n
n s = pn (m)(s + r) pn (1)s.

By inductive hypothesis, (2.2) holds for 0, . . . , i 1. We get:


(m)

i,m =

i
X

(1)j1

n+j
n

j=1

(1)i
=

i
hX

n+m+i
n




(s + r)

(1)j1

n+j
n

(1)j1

n+j
n

j=1

i
hX
j=1


p(ij)
(m)(s + r) p(ij)
(1)s +
n
n
n+i1
n

 
s =

 (ij)
pn (m) + (1)i

 (ij)
pn (1) + (1)i

i

n+m+i
n

n+i1
n

(s + r)

i
s.
(i)

The result follows from the observation that the univariate polynomial pn (k) coincides with:
i
X
j=1

(1)j1

n+j
n

 (ij)
pn (k) + (1)i

n+k+i
n


,

because both polynomials have degree n and take the value at k = 0, 1, . . . , n.

(0)

(0)
E1,2 P2

(7 6, 5) X

E >1
(0)
E >2

(0)

E >3

(1)

(1)
E2,2 P2

(14 18, 12) X

E >1

(11 11, 9) X

(1)
E >2

(21 29, 19) X

(13 15, 11) X

(20 26, 18) X

(12 13, 10) X

(19 24, 17) X

E >2

(27 34, 24) X


(27 33, 24) X
(26 31, 23) X
(26 30, 23) X

(18 21, 16) X

(21 28, 19) X

(25 28, 22) X

(18 20, 16) X

..
.

(21 29, 19) X

(17 18, 15) X

(25 27, 22) X

..
.

(16 16, 14) X

(20 26, 18) X

(29 39, 26) X


(28 36, 25) X

(19 23, 17) X

(23 34, 21) X


(22 31, 20) X

..
.

(2)

(2)
E2,3 P2

(24 26, 21) X


(24 25, 21) X

(20 25, 18) X

(23 23, 20) X

(19 23, 17) X


(19 22, 17) X
(18 21, 16) X
(18 20, 16) X
(17 18, 15) X
(1)

(0)

(1)

(1)
E1,3 P3

E >1

..
.

(15 21, 12) X

(2)

(b) E2,2 over P2 .

(a) E1,2 over P2 .

(0)

(0)
E2,3 P3

(17 26, 14) X

E >1

(c) E2,3 over P2 .


(0)

(0)
E1,4 P3

E >1

(16 23, 13) X

(14 17, 11) X

(16 22, 13) X

(17 22, 13) X

(13 15, 10)

(15 20, 12)

(16 20, 12)

(13 14, 10) X

(15 19, 12) X

(16 19, 12) X

(12 11, 9)

(15 18, 12) X

(16 18, 12) X

(14 16, 11) X

..
.

(14 15, 11) X


(13 13, 10)

(17 23, 13) X

(15 16, 11) X

..
.

(15 15, 11) X


(14 13, 10) X

(12 10, 9)

(d)

(1)
E1,3

(18 26, 14) X

(14 18, 12) X

over

P3 .

(e)

(0)
E2,3

over

P3 .

(13 10, 9)

(f)

(0)
E1,4

over P3 .

Figure 2. Examples of possible constant rank matrices arising from generalized


Steiner bundles. The notation (a b, ) stands for an a b matrix of constant rank
. Thanks to Lemma 2.2 these trees can be deduced without having to construct
the module E. For every output (a b, ) we take a random Artinian module
M with the appropriate Betti numbers, and look at the module HomR (E, M )0 .
The symbols appearing in the rightmost column indicate whether a surjective
morphism HomR (E, M )0 exists (X), does not exist (), or the computation
is too heavy to be performed (?).
2.2. Linear monads and instanton bundles. Mathematical instanton bundles were first introduced in [OS86] as rank 2m bundles on P2m+1 satisfying certain cohomological conditions. They
generalize particular rank 2 bundles on P3 whose study was motivated by problems from physics,
see [AHDM78]. They can also be defined as cohomology of a linear monad; in our work we consider
an even more general definition, in the spirit of [Jar06]. First, recall that a monad (on Pn ) is a
complex:
f

A
B
C
of vector bundles over Pn which is exact everywhere but in the middle, and such that Im f is a
subbundle of B. Its cohomology is the vector bundle E = Ker g/ Im f .
Definition 2.2. A generalized instanton bundle is a rank r vector bundle E(r,k) on Pn , r > n 1,
which is the cohomology of a linear monad of type:
(2.3)

OP2k+r

OPn (1)k .
OPn (1)k
n

10

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

When so, k = dim H1 (E(r,k) (1)) and is called the charge of E(r,k) . E(r,k) is sometimes called a
k-instanton.
According to [Fl00], the condition r > n 1 is equivalent to the existence of a monad of type
(2.3). It should be noted that it is not always the case that E(r,k) is a vector bundle; again from
[Fl00] we learn that the degeneracy locus of the map f has expected codimension r + 1. Thus
when dealing with such monads it will be necessary to check that this expected dimension indeed
corresponds to the real dimension.
Figures 3a to 3f are examples of sizes and ranks of matrices that can arise from generalized
instantons.

E(2,2) P3

E >2

E >3

(11 14, 9) X


E(2,4) P3

(19 30, 17) X


E(3,2) P3

E >2

(21 34, 18) X

(10 11, 8) X

(18 27, 16) X

(20 31, 17) X

(9 8, 7)

(18 26, 16) X

(20 30, 17) X

(29 58, 27) X

(17 24, 15) X

(28 55, 26) X

(17 23, 15) X

..
.

..
.

..
.

E >3

(17 22, 15) X

..
.

(18 22, 16)

(18 22, 15) X

..
.

(16 20, 14) X


(16 19, 14) X

(17 21, 14)


(17 20, 14) X

(18 21, 16)

(15 17, 13) X

(16 18, 13)

(17 18, 15)

(14 14, 12) X

(16 17, 13) X

(a) Classical (rank 2) instanton bun- (b) Classical (rank 2) instanton bun- (c) Generalized instanton bundle
dle over P3 of charge 2.
dle over P3 of charge 4.
E(3,2) over P3 .

E(2,2) P2

E >2

(9 9, 7) X


E(3,2) P4

E >2

(24 46, 21) X


E(4,1) P4

E >1

(24 45, 21) X


E >3

..
.

(17 23, 15) X


(16 20, 14) X

..
.

(15 18, 13) X


(15 17, 13) X
(14 15, 12) X

(d) Generalized instanton bundle


E(2,2) over P2 .

..
.
(18 22, 15)

(13 16, 7) X
(13 15, 9) X

..
.

(12 12, 8) X
(12 11, 8) X
(11 8, 7)

(18 21, 15)


(17 18, 14)

(e) Generalized instanton bundle


E(3,2) over P4 .

(f) Generalized instanton bundle


E(4,1) over P4 .

Figure 3. Examples of possible constant rank matrices that can be obtained


from instanton bundles. The notation is the same as in Figure 2.
2.3. Null correlation, Tango, and the Horrocks-Mumford bundle. Null correlation bundles
are examples of rank n 1 bundles on Pn for n odd: they are constructed as kernel of the bundle
epimorphism TPn (1) OPn (1). A construction due to Ein [Ein88] generalizes this definition on
P3 :
Definition 2.3. [Ein88] A rank 2 vector bundle E(e,d,c) on P3 is said to be a generalized null
correlation bundle if it is given as the cohomology of a monad of the form:
OP3 (c) OP3 (d) OP3 (e) OP3 (e) OP3 (d) OP3 (c),
where c > d > e > 0 are given integers.
Figures 4a to 4c show possible sizes of matrices appearing from null correlation bundles.
A construction of Tango [Tan76] produces an indecomposable rank n 1 bundle over Pn , for
all Pn , defined as a quotient En of the dual of the kernel of the evaluation map of 1Pn (2), which
is a globally generated bundle. More in detail, one starts by constructing the rank n2 bundle En
from the exact sequence:
(n+1)
(2.4)
0 TPn (2) OPn2 En 0,

11


NC P3

E >2

(15 25, 13) X


NC P5

E >1

(14 22, 12) X

(13 15, 9) X

E >2

(14 21, 12) X

E >4

(51 113, 49) X


(50 110, 48) X

(63 190, 59) X

(50 109, 48) X


(49 107, 47)

(13 18, 11) X

..
.

(13 17, 11) X

(29 29, 25) ?

(13 19, 11)

..
.


E(0,1,2) P3

(13 14, 9) X

..
.

(12 15, 10) X


(12 14, 10) X

..
.

..
.

..
.

(27 32, 25) ?


(26 30, 24) ?

(11 12, 9)

(28 26, 24) ?

(26 29, 24) ?

(10 9, 8)

(28 25, 24) ?

(25 27, 23) ?

P3 .

P5 .

(a) NC bundle over

(b) NC bundle over

(c) E(0,1,2) over P3 .

Figure 4. Size and rank of matrices of constant rank that can be constructed
from null correlation bundles. The notation is the same as in Figure 2.
and then take its quotient En :
(n)n
En En 0,
0 OPn2

(2.5)

that turns out to be a rank n indecomposable vector bundle on Pn containing a trivial subbundle
of rank 1. The Tango bundle Fn is defined as the quotient of En by its trivial subbundle, and thus
has rank n 1.
Indecomposable rank n 2 bundles on Pn are even more difficult to construct; on P4 there is
essentially only one example known, whose construction is due to Horrocks and Mumford [HM73].
It is an indecomposable rank 2 bundle that can be defined as the cohomology of the monad:
OP4 (1)5 (2P4 (2))2 OP54 .
Figures 5a to 5c show possible examples of constant rank matrices that one can construct starting
from generalized null correlation, Tango, and the Horrocks-Mumford bundle.

T P4

E >1

..
.

(29 65, 26) X


T P5

E >1

(47 129, 43) X


HM P4

E >5

(99 286, 97) X

(29 64, 26) X

(47 128, 43) X

(99 285, 97) X

(28 61, 25)

(46 124, 42)

(99 282, 97)

..
.

..
.

(99 281, 97) X

(18 21, 15)

(25 26, 21) ?

(18 20, 15)


(18 19, 15)

..
.

..
.

(25 25, 21) ?


(25 24, 21) ?

..
.
(38 38, 36) ?
(37 38, 35) ?

(17 17, 14)

(24 22, 20) ?

(37 37, 35) ?

(17 16, 14)

(24 21, 20) ?

(37 36, 35) ?

(a) Tango bundle over

P4 .

(b) Tango bundle over

P5 .

(c) Horrocks-Mumford bundle.

Figure 5. Examples of size and rank of matrices that can be constructed from
Tango and Horrocks-Mumford bundles. The notation is the same as in Figure 2.
3. Skew-symmetric matrices
We examine now the situation for linear matrices of constant rank with symmetry. The kernel
and cokernel sheaves K and E of such a matrix will be tightly related to one another, and the
matrix itself is expressed by an extension class Ext2 (E, K). Let us connect this with our map .
Definition 3.1. Let E and K be vector bundles on Pn . For t Z, consider the Yoneda map t :
t : H0 (E(t)) Ext2 (E, K) H2 (K(t)).

12

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

Set E = H0 (E) and M = H2 (K). Define as the linear map induced by the t :
: Ext2 (E, K) HomR (E, M )0 .
Theorem 3.1. Assume n > 3 and let A : R(m 1)b R(m)a be skew-symmetrizable of
constant rank. Set K = ker A and E = Coker A. Then K E (2m 1), and there is an element
lying in H2 (S 2 E (2m 1)) under the canonical decomposition
Ext2 (E, E (2m 1)) H2 (S 2 E (2m 1)) H2 (2 E (2m 1)),
such that A presents ker (). Conversely, if H2 (S 2 E (2m 1)) and = () satisfies the
assumptions of Theorem 2.1 and ker A E (2m 1), then A is skew-symmetrizable.
The same holds for symmetrizable A replacing the condition on with H2 (2 E (2m 1)).
Proof. Let us check the first statement. Assume thus that A is skew-symmetric. Then, sheafifying
the matrix A provided by Theorem 2.1 we get a long exact sequence:
A

0 K OPn (m 1)a OPn (m)a E 0

(3.1)

with K = ker A and E = Coker A. Since A has constant rank, E and K are locally free and hence:
Exti (E, OPn ) = Exti (K, OPn ) = 0,
for all i > 0. Therefore, dualizing the above sequence and twisting by OPn (2m 1)we get:
A

0 E (2m 1) OPn (m 1)a OPn (m)a K (2m 1) 0


Since the image E of A is obviously the same as the image of A, these exact sequences can be
put together to get K E (2m 1). We may thus rewrite them as:
A

0 E (2m 1) OPn (m 1)a OPn (m)a E 0.


This long exact sequence represents an element Ext2 (E, E (2m1)) H2 (E E (2m1)).
By looking at the construction of [BFM13, Lemma 3.1], it is now clear that for A to be skewsymmetric should lie in H2 (S 2 E (2m 1)).
To understand why A presents ker (), let us first expand some details of the definition of .
Let again E be the image of A and write for any integer t the exact commutative diagram:
0

K(t)

OPn (m 1 + t)a

OPn (m + t)a

E(t)

0.

E(t)
0

Taking cohomology, we get maps:


(3.2)

t : H0 (E(t))

H2 (K(t)).
H1 (E(t))

Remark that sequence (3.1) corresponds to Ext2 (E, K). Cup product with induces via
Yonedas composition the linear maps t s of (3.2). But these maps are obtained from the t by
transposition, so cup product with gives:
= t t = () : H0 (E) = E M = H2 (E (2m 1)).
This is obviously a morphism, homogeneous of degree 0. Notice that as soon as n > 3, both
groups H1 (OPn (m + t)) and H2 (OPn (m 1 + t)) vanish for all values of t, hence is surjective.
By construction A appears as presentation matrix of F = ker .
For the converse statement, the element corresponds to a length-2 extension of E (2m 1)
by E. Set = (). Theorem 2.1 gives a linear matrix A of constant rank of presentation for

13

F = ker . Note that sheafifying F we get back the bundle E, as M = H2 (E (2m 1)) is
Artinian. On the other hand, since ker A E (2m 1), the matrix A represents the extension
class . Therefore, A is skew-symmetrizable by [BFM13, Lemma 3.5 (iii)]. Part (i) of the same
lemma says that, when dealing with symmetric matrices, one should replace the condition
H2 (S 2 E (2m 1)) with H2 (2 E (2m 1)). The theorem is thus proved.


In particularly favorable situations, for instance when E is an instanton bundle of charge 2 (cf.
[BFM13, Theorem 5.2]), or a generic instanton bundle of charge 4 (cf. [BFM13, Theorem 6.1]),
one can use using results from [Dec90] and [Rah97], to check that the map is a surjection. This
makes considerably easier the search for an element corresponding to a skew-symmetric matrix
of the prescribed size and constant rank.
Example 3.1. Let us work out the case of generic instantons of rank 2 to write down an explicit
10 10 skew-symmetric matrix of constant rank 8. Let E be a rank 2 instanton bundle of charge 2
on P3 , obtained as cohomology of a monad of type (2.3). We take a special 2-istanton, as described
in [AO95]; its monad has maps:

f =

0
x1
x0
0
x3
x2

x1
x0
0
x3
x2
0

and g = x2

x3
x2

0
x3

x0
0

x1
x0

0
x1

Once the bundle E is constructed, we call E := H0 (E) its module of sections, having resolution:
0 R(4)2 R(3)6 R(2)4 R(1)2 E 0.
We truncate it in degree m = 2 in order to get a 2-linear resolution; then we take the 2nd graded
cohomology module H2 (E), which is a module of length 3, and we set M := (H2 (E)(5))>2 . The
truncated module M has length 2. By taking a morphism E >2 M we get the following diagram:
R(4)10

R(3)18

R(2)12

E >2

R(4)12

R(3)8

R(2)2

Notice that we are in a situation that is very similar to what is described in the hypotheses of
Theorem 1.1, yet here the first column cannot be surjective. Nevertheless, by resolving the kernel
of a general element 2 , we obtain a 10 10 matrix of constant rank 8, which does not enjoy any
particular symmetry property.
If we can make sure that the map E M comes indeed from an element of H2 (S 2 E (5)), then
Theorem 3.1 will guarantee that this matrix is skew-symmetrizable. By [BFM13, Theorem 5.2] we
know that H2 (S 2 E (5)) surjects onto HomR (E, H2 (E)(5))0 , and in fact an element there will
have the same kernel as its truncation in degree 2 2 , because the map is an isomorphism in degree
1. We can thus take a random element in HomR (E, H2 (E)(5))0 and our construction will work

14

ADA BORALEVI, DANIELE FAENZI, AND PAOLO LELLA

without us having to truncate. This yields the matrix A = A0 x0 + A1 x1 + A2 x2 + A3 x3 , where:


0 108 594 54 36 876 108 18 0 0

A0 =

A1 =

A2 =

and A3 =

324
0

0
0

0
0
0

162
0
0

192
192
0
18
0

18
36
0
0

0
0
0
0

64
16
16
0
0

0
0
0
18
48
0

492
48
264
24
16
0

36
18
0
0
36
36
0

0
0
0
0
0
0
0
0

193
4
41
2
163
4
49
0
89
2
17
2
0

324
0
0
0
4
48
0

0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0

27
2

438
0

534
300
0

498
0

978
612
0

108
0
54
0

319
4

23/2
35
4
0

36
18
36
0
0

64
16
16
0
0

495
2
75
705
2
27
2
18
219
0

1590
0
876
0
18
0

1058
3
368
3
2116
3
23
2

16
0

438
48
444
1
2

4
128
3
0

64
16
16
0
0
16
4
0

0
0
0
0
0
0
0
0
0
0

36
18
36
0
0
18
18
0

324
0
0
0
0
0
81
0
0

0
0
0

0
0
0
0
0
24
0
0
0
0

27
2

0
144
0
0
0

54
0
0
0
0
0
0
0
0
0

15

Acknowledgements
We would like to express our gratitude to Emilia Mezzetti for introducing us to this problem.
References
[AHDM78] M. F. Atiyah, N. J. Hitchin, V. G. Drinfel d, and Y. I. Manin, Construction of instantons, Phys. Lett.
A 65 (1978), no. 3, 185187.
[AO95]
V. Ancona and G. Ottaviani, On moduli of instanton bundles on P2n+1 , Pacific J. Math. 171 (1995),
no. 2, 343351.
[BFM13] A. Boralevi, D. Faenzi, and E. Mezzetti, Linear spaces of matrices of constant rank and instanton
bundles, Adv. Math. 248 (2013), 895920.
[BM14]
A. Boralevi and E. Mezzetti, Planes of matrices of constant rank and globally generated vector bundles,
to appear in Ann. Inst. Fourier (Grenoble), 2014.
[Dec90]
W. Decker, Monads and cohomology modules of rank 2 vector bundles, Compositio Math. 76 (1990),
no. 12, 717.
[EG84]
D. Eisenbud and S. Goto, Linear free resolutions and minimal multiplicity, J. Algebra 88 (1984), no. 1,
89133.
[EH88]
D. Eisenbud and J. Harris, Vector spaces of matrices of low rank, Adv. in Math. 70 (1988), no. 2,
135155.
[Ein88]
L. Ein, Generalized null correlation bundles, Nagoya Math. J. 111 (1988), 1324.
[Fl00]
G. Flystad, Monads on projective spaces, Comm. Algebra 28 (2000), no. 12, 55035516, Special issue
in honor of Robin Hartshorne.
[HM73]
G. Horrocks and D. Mumford, A rank 2 vector bundle on P 4 with 15,000 symmetries, Topology 12
(1973), 6381. MR 0382279
[IL99]
B. Ilic and J.M. Landsberg, On symmetric degeneracy loci, spaces of symmetric matrices of constant
rank and dual varieties, Math. Ann. 314 (1999), no. 1, 159174.
[Jar06]
M. Jardim, Instanton sheaves on complex projective spaces, Collect. Math. 57 (2006), no. 1, 6991.
[OS86]
C. Okonek and H. Spindler, Mathematical instanton bundles on P2n+1 , J. Reine Angew. Math. 364
(1986), 3550.
[Pee11]
I. Peeva, Graded syzygies, Algebra and Applications, vol. 14, Springer-Verlag London, Ltd., London,
2011.
[Rah97]
O. Rahavandrainy, R
esolution des fibr
es instantons g
en
eraux, C. R. Acad. Sci. Paris S
er. I Math. 325
(1997), no. 2, 189192.
[Syl86]
J. Sylvester, On the dimension of spaces of linear transformations satisfying rank conditions, Linear
Algebra Appl. 78 (1986), 110.
[Tan76]
H. Tango, An example of indecomposable vector bundle of rank n 1 on P n , J. Math. Kyoto Univ. 16
(1976), no. 1, 137141.
[Wes96]
R. Westwick, Spaces of matrices of fixed rank. II, Linear Algebra Appl. 235 (1996), 163169.
TU Eindhoven, Department of Mathematics and Computer Science, PO Box 513, 5600 MB Eindhoven,
The Netherlands
E-mail address: a.boralevi@tue.nl
e de Bourgogne, 9 Avenue Alain
Institut de Mathmatiques de Bourgogne, UMR CNRS 5584, Universit
Savary, BP 47870, 21078 Dijon Cedex, France
E-mail address: daniele.faenzi@u-bourgogne.fr
` degli Studi di Trento, Via Sommarive 14, 38123 Trento,
Dipartimento di Matematica, Universita
Italy
E-mail address: paolo.lella@unitn.it
URL: http://www.paololella.it/

You might also like