on varieties of commuting nilpotent matrices

26
Linear Algebra and its Applications 452 (2014) 237–262 Contents lists available at ScienceDirect Linear Algebra and its Applications www.elsevier.com/locate/laa On varieties of commuting nilpotent matrices Nham V. Ngo a , Klemen Šivic b,a Department of Mathematics and Statistics, Lancaster University, Lancaster, LA1 4YW, United Kingdom b Department of Mathematics, ETH Zürich, Rämistrasse 101, Zürich, Switzerland article info abstract Article history: Received 5 June 2013 Accepted 22 March 2014 Available online 16 April 2014 Submitted by R. Guralnick MSC: 15A27 Keywords: Irreducibility of varieties of commuting nilpotent matrices Approximation by 1-regular nilpotent matrices Let N (d, n) be the variety of all d-tuples of commuting nilpotent n × n matrices. It is well-known that N (d, n) is irreducible if d = 2, if n 3 or if d = 3 and n = 4. On the other hand N (3,n) is known to be reducible for n 13. We study in this paper the reducibility of N (d, n) for various values of d and n. In particular, we prove that N (d, n) is reducible for all d, n 4. In the case d = 3, we show that it is irreducible for n 6. © 2014 Elsevier Inc. All rights reserved. 1. Introduction Let C (d, n) denote the set of all d-tuples of commuting n × n matrices over an alge- braically closed field F and let N (d, n) be its subset consisting of all d-tuples of nilpotent commuting matrices. Both sets are defined by polynomial equations in the entries of ma- trices, therefore they can be viewed as affine varieties in F dn 2 . Irreducibility of C (d, n) has been studied for a long time. Motzkin and Taussky [15] proved that the variety C (2,n) is irreducible for each positive integer n. On the other hand, C (d, n) is known * Corresponding author. E-mail addresses: [email protected] (N.V. Ngo), [email protected] (K. Šivic). http://dx.doi.org/10.1016/j.laa.2014.03.032 0024-3795/© 2014 Elsevier Inc. All rights reserved.

Upload: klemen

Post on 30-Dec-2016

225 views

Category:

Documents


4 download

TRANSCRIPT

Linear Algebra and its Applications 452 (2014) 237–262

Contents lists available at ScienceDirect

Linear Algebra and its Applications

www.elsevier.com/locate/laa

On varieties of commuting nilpotent matrices

Nham V. Ngo a, Klemen Šivic b,∗

a Department of Mathematics and Statistics, Lancaster University, Lancaster,LA1 4YW, United Kingdomb Department of Mathematics, ETH Zürich, Rämistrasse 101, Zürich, Switzerland

a r t i c l e i n f o a b s t r a c t

Article history:Received 5 June 2013Accepted 22 March 2014Available online 16 April 2014Submitted by R. Guralnick

MSC:15A27

Keywords:Irreducibility of varieties ofcommuting nilpotent matricesApproximation by 1-regularnilpotent matrices

Let N(d, n) be the variety of all d-tuples of commutingnilpotent n × n matrices. It is well-known that N(d, n) isirreducible if d = 2, if n � 3 or if d = 3 and n = 4. Onthe other hand N(3, n) is known to be reducible for n � 13.We study in this paper the reducibility of N(d, n) for variousvalues of d and n. In particular, we prove that N(d, n) isreducible for all d, n � 4. In the case d = 3, we show that it isirreducible for n � 6.

© 2014 Elsevier Inc. All rights reserved.

1. Introduction

Let C(d, n) denote the set of all d-tuples of commuting n × n matrices over an alge-braically closed field F and let N(d, n) be its subset consisting of all d-tuples of nilpotentcommuting matrices. Both sets are defined by polynomial equations in the entries of ma-trices, therefore they can be viewed as affine varieties in F

dn2 . Irreducibility of C(d, n)has been studied for a long time. Motzkin and Taussky [15] proved that the varietyC(2, n) is irreducible for each positive integer n. On the other hand, C(d, n) is known

* Corresponding author.E-mail addresses: [email protected] (N.V. Ngo), [email protected] (K. Šivic).

http://dx.doi.org/10.1016/j.laa.2014.03.0320024-3795/© 2014 Elsevier Inc. All rights reserved.

238 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

to be reducible if d and n are both at least 4. This result is usually referred to Gersten-haber [5], but he proved reducibility of C(d, n) only if n � 4 and d � n + 1 (and hisproof can be applied also to d = n � 4), while reducibility of C(d, n) for d � 4 and n � 4was proved by Guralnick [6], using Gerstenhaber’s idea. It is also known that C(d, n)is irreducible for each d if n � 3 (see [12] and [6]). The situation in the case of triplesis much more complicated, and the problem of irreducibility of C(3, n) is not solvedcompletely. With a dimension argument Guralnick [6] proved that C(3, n) is reduciblefor n � 32, and using his idea Holbrook and Omladič showed reducibility of C(3, n) forn � 30. In fact, their proof in [11] shows also reducibility of C(3, 29), since in the caseof irreducibility of C(3, n) all its subvarieties must have strictly smaller dimension thann2+2n, and the inequality at the beginning of page 136 of [11] does not need to be strict.On the other hand, C(3, n) is known to be irreducible for n � 10 (see [8,11,20,9,24,26]).The last results were stated if charF = 0, but in many proofs this assumption can beomitted, and probably slight modifications of the other proofs would also give the sameresults in any characteristic.

In this paper we study reducibility of N(d, n) and the characteristic of F will bearbitrary. The study of N(d, n) started much later than that of C(d, n). Baranovsky [1]and Basili [3] independently showed that N(2, n) is irreducible for each n, and this resultwas significantly generalized by Premet [21] to varieties of pairs of commuting nilpotentelements of arbitrary reductive Lie algebras. Explicitly, he proved that these varietiesare equidimensional and their components are parametrized by distinguished nilpotentconjugacy classes in the Lie algebra. Moreover, his proof for irreducibility of N(2, n) doesnot depend on the characteristic of F, while in [1] and [3] there are some restrictions oncharF. On the other hand, for d � 3 much less is known. If n � 3, then N(d, n) isirreducible for each d by Proposition 5.1.1 and Theorem 7.1.2 of [18]. In the introductionof Chapter II of Young’s thesis [28] it was claimed that N(d, n) is reducible for d � 4and n � 4. However, this result was not proved and it is also not clear how to obtainit. To show reducibility of C(d, n) for n � 4 and d � n Gerstenhaber [5] constructeda unital algebra (i.e. an algebra with an identity) of dimension more than n which isgenerated by d commuting (square zero) n × n matrices, and his proof can clearly beapplied to the nilpotent case to show reducibility of N(d, n) for d � n � 4. However,Guralnick’s [6] proof of reducibility of C(d, n) for n � 4 and d � 4, using block diagonalmatrices, cannot be applied to the nilpotent case, since all blocks have to be nilpotent.Using Gerstenhaber’s idea we will prove that N(d, n) is indeed reducible for all d � 4 andn � 4. In particular, for each n � 4 we will find a unital (n+1)-dimensional commutativesubalgebra of Mn(F) generated by 4 nilpotent matrices. The question of irreducibility ofN(d, n) then remains open only in the case of triples, i.e. for d = 3. There are two knownresults about irreducibility of N(3, n). In [28, Theorem 2.2.1] it was proved that N(3, 4) isirreducible. On the other hand, Clark, O’Meara and Vinsonhaler proved in their book onWeyr form [19] that the variety N(3, n) is reducible for n � 13 (Theorem 7.10.5), usingmethods from [6] and [11]. Since this result was proved for much smaller dimensionthan analogous result for C(3, n), it seems to be easier to determine when N(3, n) is

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 239

irreducible than when C(3, n) is. In this paper we partially answer this question. Weprove irreducibility of N(3, n) for n � 6. Our result provides a rigorous argument forirreducibility of N(3, 4) stated in Young’s thesis [28]. We also prove that the variety ofpairs of commuting nilpotent matrices in the centralizer of a 2-regular nilpotent matrix(i.e. matrix whose Jordan canonical form has at most 2 Jordan blocks) is irreducible,which gives an analog to Corollary 10 of [17]. However, we will show that the same resultdoes not hold for 3-regular case, i.e. a result analogous to Theorem 12 of [25] does nothold in the nilpotent case.

The paper is organized as follows. In Section 2 we recall some preliminary results onnilpotent commuting matrices from the literature. We show that the variety N(d, n) isirreducible if and only if it is equal to the Zariski closure of the set of all d-tuples ofcommuting nilpotent 1-regular n × n matrices. As a corollary we show two interestingrelations between (ir)reducibility of varieties N(d, n) and C(d, n). In Section 3 we showthat for each n � 4 there exists a unital (n + 1)-dimensional commutative subalgebraof Mn(F) which is generated by 4 nilpotent matrices. This will imply reducibility ofvarieties N(d, n) for all d � 4 and n � 4. In Section 4 we consider varieties of pairs ofcommuting nilpotent matrices in the centralizers of given nilpotent matrices. We showthat such variety is irreducible if the given matrix is 2-regular, but it can be reducible ifthe given matrix is 3-regular. In Section 5 we prove irreducibility of varieties N(3, 5) andN(3, 6), using simultaneous commutative perturbations of triples of commuting nilpotentmatrices by triples of commuting nilpotent 1-regular matrices. We also show that inany dimension a triple of nilpotent commuting matrices can be perturbed by triples of1-regular nilpotent commuting matrices if Jordan canonical forms of the matrices in thetriple have only one nonzero Jordan block or all Jordan blocks of order at most 2. Notethat analogous results in the case of C(3, n) were proved in [11].

2. Preliminaries

In the proofs of (ir)reducibility of C(d, n) generic matrices play an important role.Recall that an n × n matrix is called generic if it has n distinct eigenvalues. Genericmatrices form an open subset of Mn(F) defined by the inequality det p′X(X) �= 0 wherep′X is the derivative of the characteristic polynomial of the matrix X. In particular, eachirreducible component of the subvariety of Mn(F) consisting of all matrices that arenot generic is (n2 − 1)-dimensional. Moreover, the variety C(d, n) is irreducible if andonly if it is equal to the Zariski closure of the set of all d-tuples of generic commutingn× n matrices (see [26], and note that these results do not depend on the characteristicof F). To adapt this strategy for proving irreducibility of N(d, n), we use the conceptof 1-regular matrices instead. In this section we recall some results from the literature,especially from [19], on commuting nilpotent matrices, that will be needed in the sequel.We start with the following definition from [17].

240 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

Definition. A matrix A ∈ Mn(F) is called r-regular if each its eigenspace is at mostr-dimensional.

Remark. Note that the eigenspace of a nilpotent matrix is its kernel and that a nilpotentmatrix is r-regular if and only if its Jordan canonical form has at most r Jordan blocks.

For the study of commutativity 1-regular matrices (which are also called regular ornonderogatory) are especially important, since each matrix that commutes with 1-regularmatrix is a polynomial in that matrix. Moreover, each (nilpotent) n × n matrix canbe perturbed by (nilpotent) 1-regular matrices, and therefore it belongs to the Zariskiclosure of the set of (nilpotent) 1-regular matrices. We denote by Nn the variety of allnilpotent n×n matrices, by Rn the set of all 1-regular nilpotent n×n matrices, by R(d, n)the set of all d-tuples (A1, A2, . . . , Ad) ∈ N(d, n) such that the linear span of A1, . . . , Ad

contains a 1-regular matrix, and by R1(d, n) the set of all d-tuples (A1, . . . , Ad) ∈ N(d, n)with A1 1-regular. By Proposition 1 of [17] the set of all 1-regular n×n matrices is openin Mn(F) in the Zariski topology, therefore Rn is open in Nn and R1(d, n) is open inN(d, n). On the other hand, the set R(d, n) is obtained by the action of the general lineargroup GLd(F) on R1(d, n), hence it can be written as a union of the sets isomorphic toR1(d, n). Therefore R(d, n) is also open in N(d, n). In particular, if N(d, n) is irreducible,then R(d, n) = R1(d, n) = N(d, n), where S denotes the closure of the set S in the Zariskitopology. We first show that the converse of this implication also holds. If follows fromthe following proposition (which is essentially Lemma 7.10.3 of [19]) that shows theirreducibility and the dimension of R1(d, n).

Proposition 1. The variety R1(d, n) is irreducible and of dimension (n + d− 1)(n− 1).

Proof. For any nonnegative integer k we denote by Fk[t] the space of all polynomialsin t over F of degree at most k. The polynomial map ϕ :Rn × (Fn−2[t])d−1 → R1(d, n)defined by

ϕ(A, p1, p2, . . . , pd−1) =(A,Ap1(A), Ap2(A), . . . , Apd−1(A)

)is bijective, since all matrices that commute with 1-regular nilpotent matrix A are poly-nomials in A, and the constant terms of these polynomials have to be zero because ofthe nilpotency of the matrices in a d-tuple from N(d, n). The variety Nn = Rn is ir-reducible and of dimension n2 − n by Proposition 2.1 of [3], therefore R1(d, n) is theclosure of a polynomial image of the irreducible variety Nn × (Fn−2[t])d−1, hence it isirreducible. Moreover, the theorem on fibres (Theorem 11.12 of [10]) implies that thevarieties Nn × (Fn−2[t])d−1 and R1(d, n) have the same dimensions, i.e.

dimR1(d, n) = dimNn + dim((Fn−2[t]

)d−1) = (n + d− 1)(n− 1). �

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 241

Remark. Based on this result and the arguments from [6] and [11] the reducibility ofN(3, n) for n � 13 was proved in Theorem 7.10.5 of [19].

Corollary 2. The variety N(d, n) is irreducible if and only if R(d, n) = N(d, n).

Proof. Since the variety R1(d, n) is irreducible and the set R(d, n) is obtained fromR1(d, n) by the action of the group GLd(F), the variety R(d, n) is also irreducible. More-over, since the set R1(d, n) is its open subset, the varieties R(d, n) and R1(d, n) are equal.The corollary now follows immediately from the previous proposition. �

The following two corollaries show the relations between (ir)reducibility of C(d, n)and of N(d, n).

Corollary 3. Suppose that the variety C(d,m) is irreducible for each m < n and thatN(d, n) is irreducible. Then C(d, n) is also irreducible.1

Proof. We have to show that R1(d, n) = C(d, n), where R1(d, n) is the set of all d-tuples(A1, . . . , Ad) ∈ C(d, n) where at least one of the matrices A1, . . . , Ad is 1-regular (see[25]). Let (A1, . . . , Ad) ∈ C(d, n) be an arbitrary d-tuple. If one of the matrices A1, . . . , Ad

has two distinct eigenvalues, then extending the proof of Lemma 5 of [26] (which clearlyholds in any characteristic) to any number of matrices we get (A1, . . . , Ad) ∈ R1(d, n).In the sequel we can therefore assume that Ai has only one eigenvalue λi for eachi = 1, . . . , d. If we denote by I the identity matrix, then (A1 − λ1I, . . . , Ad − λdI) ∈N(d, n) = R1(d, n) ⊆ R1(d, n) by the assumption on irreducibility of the variety N(d, n).However, the sets R1(d, n) and {(X1 +λ1I, . . . ,Xd +λdI); (X1, . . . , Xd) ∈ R1(d, n)} areclearly the same, so their closures also are, and we get (A1, . . . , Ad) ∈ R1(d, n). �Corollary 4. If the variety N(d,m) is irreducible for each m � n, then the variety C(d, n)is also irreducible.

Since we know that C(4, 4) is reducible and N(4, n) is irreducible for n � 3, we obtain:

Corollary 5. The variety N(4, 4) is reducible.

We conclude this section with two lemmas telling us which reductions we can do whileproving that N(d, n) = R(d, n). The first lemma is a simple consequence of the fact thatin an irreducible variety any nonempty open subset is dense.

Lemma 6. If V ⊆ N(d, n) is any irreducible variety and U ⊆ V any nonempty open subsetsuch that U ⊆ R(d, n), then V ⊆ R(d, n).

1 This is originally Theorem 3.2.2 in the unpublished paper [7]. The first author thanks Robert Guralnickfor sharing the idea with him.

242 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

The lemma tells us that if we want to prove that any d-tuple from some irreduciblesubvariety of N(d, n) belongs to R(d, n), then we can assume any nonempty open con-dition on the d-tuple.

Lemma 7. Let (A1, . . . , Ad) ∈ N(d, n) be a d-tuple of commuting nilpotent matrices,P ∈ GLn(F) an arbitrary invertible matrix, Q ∈ GLn(F) an invertible matrix satisfyingQ−1AT

1 Q = A1, and p2, . . . , pd ∈ F[t] arbitrary polynomials with zero constant terms.Then (A1, . . . , Ad) ∈ R(d, n) if and only if any of the following holds.

(a) (P−1A1P, . . . , P−1AdP ) ∈ R(d, n).

(b) (B1, . . . , Bd) ∈ R(d, n) where (B1, . . . , Bd) ∈ N(d, n) is any d-tuple of nilpotentcommuting matrices such that the linear spans of A1, . . . , Ad and of B1, . . . , Bd arethe same.

(c) (A1, A2 − p2(A1), . . . , Ad − pd(A1)) ∈ R(d, n).(d) (AT

1 , . . . , ATd ) ∈ R(d, n).

(e) (A1, Q−1AT

2 Q, . . . , Q−1ATd Q) ∈ R(d, n).

Proof. Let ϕ :N(d, n) → N(d, n) be a polynomial map that maps the set R(d, n) toits closure. Since polynomial maps are continuous in the Zariski topology, it followsthat ϕ(R(d, n)) ⊆ ϕ(R(d, n)) ⊆ R(d, n). Moreover, if ϕ is invertible and its inverseis also a polynomial map that maps the set R(d, n) to its closure, then ϕ(R(d, n)) =R(d, n). Now the equivalences in (a), (b), (d) and (e) follow immediately if the map ϕ isrespectively conjugation of the matrices in the d-tuple by P , the action of a fixed elementof GLd(F) on the d-tuple that maps (A1, . . . , Ad) to (B1, . . . , Bd), the transposition, andthe composition of the conjugation and the transposition. However, since R1(d, n) =R(d, n), the conclusion ϕ(R(d, n)) = R(d, n) follows also if ϕ is bijective with polynomialinverse and ϕ and ϕ−1 map R1(d, n) to its closure. The equivalence in (c) then followsimmediately if ϕ(X1, X2, . . . , Xd) = (X1, X2 − p2(X1), . . . , Xd − pd(X1)). �3. Commutative subalgebras of Mn(FFF) and reducibility of N(d, n) for d, n ��� 4

By Corollary 5 the variety N(4, 4) is reducible. In this section we generalize this re-sult. In other words, we show that N(d, n) is reducible for all d, n � 4. To do so weuse the connection between irreducibility of N(d, n) and the dimension of commuta-tive subalgebras of Mn(F) generated by d nilpotent matrices. For commuting matricesA1, . . . , Ad ∈ Mn(F) let F[A1, . . . , Ad] denote the unital algebra generated by the matricesA1, . . . , Ad. Using the idea from [6] we first show that irreducibility of the variety N(d, n)implies that each commutative unital subalgebra of Mn(F) generated by d nilpotent ma-trices is at most n-dimensional. Then for each n � 4 we construct an (n+1)-dimensionalcommutative unital subalgebra of Mn(F) that is generated by 4 nilpotent matrices, whichshows reducibility of N(d, n) for all d, n � 4. We start with the following proposition

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 243

whose proof is just extension of the proof of Theorem 1 of [6] to an arbitrary number ofmatrices.

Proposition 8. For each positive integer r the set{(A1, . . . , Ad) ∈ C(d, n); dimF[A1, . . . , Ad] � r

}is closed in the Zariski topology (i.e. it is an affine variety).

Proof. Consider the n2 × nd matrix whose columns are the matrices Ak11 · · ·Akd

d consid-ered as column vectors of size n2, where 0 � ki � n− 1 for i = 1, . . . , d. The rank of thismatrix is equal to the dimension of the algebra F[A1, . . . , Ad]. However, the conditionthat the rank of a matrix is bounded by r is a closed condition, hence the propositionfollows. �Corollary 9. If the variety N(d, n) is irreducible, then each commutative unital subalgebraof Mn(F) generated by d nilpotent matrices is at most n-dimensional.

Proof. By the previous proposition the set

V ={(A1, . . . , Ad) ∈ N(d, n); dimF[A1, . . . , Ad] � n

}is a subvariety of N(d, n). It contains the set R1(d, n), since if A1 is 1-regular andA2, . . . , Ad commute with A1, then they are polynomials in A1, and F[A1, . . . , Ad] =F[A1]. Therefore it contains also the closure R1(d, n) which is equal to N(d, n), sinceN(d, n) is by the assumption irreducible. �

Now we can prove reducibility of N(d, n) if n and d are both at least 4.

Theorem 10. For all positive integers d, n � 4 the variety N(d, n) is reducible.

Proof. By the previous corollary it suffices to find for each n � 4 some commutativeunital subalgebra of Mn(F) of dimension n+1 which is generated by 4 nilpotent matrices.Let A1 be the nilpotent matrix in Jordan canonical form which has one Jordan block oforder n − 2 and one Jordan block of order 2, and let A2 = e1eTn , A3 = en−1eTn−2 andA4 = en−1eTn where ei denotes the i-th standard basis vector of Fn, i.e. the vector with 1on the i-th component and 0 elsewhere. Then the algebra F[A1, A2, A3, A4] is generatedby 4 nilpotent commuting matrices and spanned by I, A1, . . . , A

n−31 , A2, A3, A4 as a

vector space (i.e. dimF[A1, A2, A3, A4] = n + 1), which proves the theorem. �4. Pairs of commuting nilpotent matrices in the centralizer of a nilpotent matrix

In Section 2 we have seen that N(3, n) is irreducible if and only if it is equal tothe Zariski closure of R(3, n). The proof of irreducibility of N(3, n) is usually based

244 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

on perturbation by triples of 1-regular nilpotent commuting matrices, in the sense thatthrough each triple from N(3, n) we find an irreducible curve (typically a line) in N(3, n)with a Zariski open (and dense) subset contained in R(3, n). However, in general the per-turbed triples are not easy to find. Much easier to find, when they exist, are commutativenilpotent 1-regular perturbations of only two of the matrices of the triple in the central-izer of the third matrix. (For an example when this is not possible, see Proposition 23below.) Similarly as in the proofs of irreducibility of C(3, n), it is therefore convenientto study varieties of pairs of commuting nilpotent matrices in the centralizers of givennilpotent matrices. For a nilpotent matrix A ∈ Mn(F) denote by C(A) the centralizerof A, by N(A) the nilpotent centralizer of A (i.e. the set of all nilpotent n× n matricesthat commute with A), and let

C2(A) ={(B,C) ∈ C(A) × C(A); BC = CB

}and

N2(A) ={(B,C) ∈ C2(A); B and C are nilpotent

}.

Clearly all these sets are affine varieties. The variety C2(A) is irreducible if A is 3-regular,cf. Corollary 10 of [17] and Theorem 12 of [25]. In this section we study analogous problemfor N2(A). In particular we prove that the variety N2(A) is irreducible if A is 2-regularnilpotent matrix. However, the same does not hold if A is 3-regular. For the proofs ofthese results we now introduce some additional notation. For a nilpotent n × n matrixA we define

D(A) ={B ∈ N(A); dimF[A,B] = n

}and

D2(A) ={(B,C) ∈ N2(A); dimF[A,B] = n

}.

Note that, by Theorem 1.1 of [16], for commuting n×n matrices A and B the conditiondimF[A,B] = n is equivalent to dim(C(A) ∩ C(B)) = n and to the condition that thealgebra F[A,B] is self-centralizing.

Lemma 11. For any nilpotent matrix A ∈ Mn(F) the sets D(A) and D2(A) are Zariskiopen in N(A) and N2(A), respectively.

Proof. By Proposition 8 the set {(A′, B) ∈ C(2, n); dimF[A′, B] � r} is Zariski closedfor each positive integer r. On the other hand, by Theorem 1 of [6] a unital algebragenerated by two commuting n × n matrices can be at most n-dimensional, so the setV = {(A′, B) ∈ C(2, n); dimF[A′, B] �= n} is closed, and therefore ({A} ×N(A)) ∩ V is

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 245

closed. However, the later set is isomorphic to N(A)\D(A), which shows that D(A) isopen in N(A).

For the second part of the lemma observe that

N2(A)\D2(A) =((N(A)\D(A)

)×N(A)

)∩N(2, n),

which is closed, so D2(A) is open in N2(A). �The following lemma was proved in [3] (Lemma 2.3).

Lemma 12 (Basili). For any nilpotent matrix A the variety N(A) is irreducible.

Corollary 13. D(A) = N(A) for each nilpotent n× n matrix A.

Proof. Because of the previous two lemmas we have to prove only that the set D(A)is nonempty. Since for each invertible matrix P ∈ GLn(F) the varieties N(A) andN(P−1AP ) are clearly isomorphic, we can assume that the matrix A is in the Jor-

dan canonical form, i.e. A =[

J1J2

...Jk

]where Ji is the ni × ni nilpotent Jordan

block for each i = 1, 2, . . . , k, and n1 � n2 � · · · � nk � 1 (and Ji = 0 in the case of

ni = 1). Let B =[ 0 K1

0 ...... Kk−1

0

]where for each i = 1, . . . , k − 1, Ki is the ni × ni+1

matrix of the form

⎡⎣ 11

...1

⎤⎦. Then the matrix B is nilpotent, it commutes with A and

dim(kerA ∩ kerB) = 1, where kerX denotes the kernel of the matrix X. Theorem 2 of[13] then implies that dimF[A,B] = n, i.e. B ∈ D(A). Note that, although in [13] it isassumed F = C, the proof of Theorem 2 of [13] is valid over any field. �Proposition 14. For each nilpotent n× n matrix A the variety D2(A) is irreducible andof dimension dimC(A) − dim kerA + n− 1.

Proof. Similarly as in the previous corollary we can assume that the matrix A is in theJordan canonical form. Let A = J1 ⊕ · · · ⊕ Jk where for each i = 1, . . . , k the matrix Jiis the nilpotent Jordan block of size ni × ni, and n1 � n2 � · · · � nk � 1. Let B be anynilpotent n×n matrix that commutes with A such that the algebra generated by A andB is n-dimensional. Then the algebra F[A,B] has the basis

B ={AiBj ; 0 � j � k − 1, 0 � i � nj+1 − 1

}(see Theorem 2 of [2] or Theorem 1.1 of [14]). We define the polynomial map

ϕ :D(A) × Fn1−1 × F

n2 × · · · × Fnk → D2(A)

246 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

by

ϕ(B, (c21, . . . , cn11), (c12, . . . , cn22), . . . , (c1k, . . . , cnkk)

)=(B,

n1∑i=2

ci1Ai−1 +

k∑j=2

nj∑i=1

cijAi−1Bj−1

).

If (B,C) ∈ D2(A), then dimF[A,B] = n, and Theorem 1.1 of [16] implies that thealgebra F[A,B] is self-centralizing. In particular, C is a polynomial in A and B, andsince the matrix C is nilpotent and B is the basis of F[A,B], it follows that the mapϕ is surjective. Therefore the closure D2(A) is a closure of a polynomial image of theirreducible variety N(A)×F

n−1, hence it is irreducible. Moreover, since B is the basis ofF[A,B], the map ϕ is also injective, and dimD2(A) = dimN(A) +n− 1 by the theoremon fibres.

To complete the proof we have to find the dimension of the variety N(A). We can

conjugate the matrix A, so we can assume that A =

⎡⎣ J1

J2...

Jl

⎤⎦ for some positive

integer l where for each i = 1, . . . , l the matrix Ji =[ 0 I

0 ...... I

0

]has mi block rows and

columns of dimension si, and m1 > m2 > · · · > ml � 1 (where Jl = 0 if ml = 1).It is clear that a matrix B commutes with A if and only if it is of the form B =[

B11 B12 ··· B1lB21 B22 ··· B2l...

......

...Bl1 Bl2 ··· Bll

]where the matrices Bij are block upper triangular and Toeplitz, i.e.

• Bij =

⎡⎢⎢⎢⎢⎢⎢⎣B

(1)ij B

(2)ij ··· B

(mj)ij

0 B(1)ij

......

...... B

(2)ij

...... B

(1)ij

0...

0 ··· 0

⎤⎥⎥⎥⎥⎥⎥⎦ for some B(1)ij , . . . , B

(mj)ij ∈ Msi×sj (F) if i < j,

• Bij =

⎡⎢⎢⎣0 ··· 0 B

(1)ij B

(2)ij ··· B

(mi)ij

0 ··· 0 B(1)ij

......

......

... B(2)ij

0 ··· 0 B(1)ij

⎤⎥⎥⎦ for some matrices B(1)ij , . . . , B

(mi)ij ∈ Msi×sj (F) if

i > j, and

• Bii =

⎡⎢⎣B(1)ii B

(2)ii ··· B

(mi)ii

B(1)ii

......

... B(2)ii

B(1)ii

⎤⎥⎦ for some B(1)ii , . . . , B

(mi)ii ∈ Msi(F).

By Lemma 2.3 of [3] the matrix B belongs to N(A) (i.e. it is nilpotent) if and only ifthe matrices B

(1)11 , . . . , B

(1)ll are all nilpotent. Since dimNsi = dimMsi(F) − si for each

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 247

i = 1, . . . , l, we obtain dimN(A) = dimC(A) −∑l

i=1 si = dimC(A) − dim kerA, whichcompletes the proof of the proposition. �Corollary 15. Let A be a nilpotent n × n matrix. Then the variety N2(A) is irreducibleif and only if N2(A) = D2(A).

The next lemma tells us which reduction we can make when proving that a pair fromN2(A) belongs to D2(A).

Lemma 16. Let (B,C) ∈ N2(A) and let p ∈ F[t, u] be a polynomial in two variables withzero constant term. Then the pair (B,C) belongs to D2(A) if and only if any of thefollowing holds.

(a) (C,B) ∈ D2(A).(b) (B, p(A,B) + C) ∈ D2(A).

Proof. If ϕ :N2(A) → N2(A) is a polynomial map that maps D2(A) to D2(A), thenϕ(D2(A)) ⊆ ϕ(D2(A)) ⊆ D2(A), since polynomial maps are continuous in the Zariskitopology. Moreover, if ϕ is bijective and its inverse is also a polynomial map that mapsD2(A) to its closure, then ϕ(D2(A)) = D2(A). Now the second equivalence immediatelyfollows, while for the first one we have to prove that the polynomial map ϕ :N2(A) →N2(A) defined by ϕ(B,C) = (C,B) maps D2(A) to D2(A). However, this is clear, sinceD2(A) is open subset of N2(A), and if (B,C) ∈ D2(A), then (C + λB,B) ∈ D2(A) forall except finitely many scalars λ ∈ F, and therefore (C,B) ∈ D2(A). �

We can now prove the main theorem of this section. Note that Corollary 10 of [17] isthe analogous result for C2(A) instead of N2(A).

Theorem 17. For any 2-regular nilpotent n×n matrix A the variety N2(A) is irreducible.

Proof. Let (B,C) ∈ N2(A) be an arbitrary pair. We have to prove that (B,C) ∈ D2(A).If C is a polynomial in A and B or if B is a polynomial in A and C, then by Lemma 16(a)we can assume the first case, and moreover, by Lemma 16(b) we can assume that C = 0.Since the set D(A) is nonempty, there exists a nilpotent matrix X ∈ C(A) such thatdimF[A,X] = n, i.e. (X, 0) ∈ D2(A). The line {(λX+(1−λ)B, 0); λ ∈ F} then intersectsthe open subset D2(A) of N2(A), therefore (λX + (1 − λ)B, 0) ∈ D2(A) for all exceptfinitely many scalars λ ∈ F. However, then the whole line belongs to D2(A), and inparticular (B, 0) ∈ D2(A).

In the sequel we assume that C is not a polynomial in A and B and also that B isnot a polynomial in A and C. For any invertible matrix P ∈ GLn(F) the varieties N2(A)and N2(P−1AP ) are clearly isomorphic, therefore we can assume that the matrix A isin the Jordan canonical form, i.e. A =

[Jk 00 Jm

]where Jk and Jm are Jordan blocks of

orders k and m, respectively, and k � m. We will consider two cases.

248 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

Case 1. Assume that k > m. By Lemma 4 of [17] the elements of the centralizerC(A) can be identified with the matrices of the form

[p(t) tk−mq(t)r(t) s(t)

]where the first row

belongs to F[t]/tk and the second row belongs to F[t]/tm, i.e. p, q ∈ F[t]/tk, r, s ∈ F[t]/tmand q is of degree less than m. Moreover, as observed in the remark after Lemma 4of [17] the multiplication in C(A) corresponds to the multiplication of such polynomialmatrices. In Case 1 we will use this polynomial notation for all elements of C(A). Notethat in this notation A corresponds to the matrix

[t 00 t

]. The matrices B and C commute

with A, they are nilpotent and by Lemma 16 we can add any polynomials (with zeroconstant terms) in A to them, therefore we can assume that B =

[0 tk−mp(t)

q(t) r(t)

]and

C =[

0 tk−mp′(t)q′(t) r′(t)

]for some polynomials p, p′, q, q′, r, r′ ∈ F[t]/tm. Since B and C are

nilpotent, the polynomials r and r′ have zero constant terms.(1) Assume first that at least one of the polynomials p, p′, q, q′ has nonzero con-

stant term. By Lemma 16(a) we can assume that p or q has nonzero constant term.Then it is easy to see that in the first case dim(kerA ∩ kerB) = 1 and in the sec-ond case dim(kerAT ∩ kerBT ) = 1. Theorem 2 of [13] then in both cases implies thatdimF[A,B] = n, therefore (B,C) ∈ D2(A).

(2) Assume now that the constant terms of the polynomials p, p′, q, q′ are all zero.Let β be the largest positive integer such that tβ divides p(t), q(t) and r(t), and letγ be the largest positive integer such that tγ divides p′(t), q′(t) and r′(t). Then B =Aβ[

0 tk−mp(t)q(t) r(t)

]and C = Aγ

[0 tk−mp′(t)

q′(t) r′(t)

], where at least one of the polynomials

p(t), q(t), r(t) is not divisible by t and at least one of p′(t), q′(t), r′(t) is not divisible by t.If t divides r(t) or r′(t), then by Lemma 16(a) we can assume that t divides r(t). Thematrix X =

[0 tk−mp(t)

q(t) r(t)

]is then nilpotent and it commutes with B. Since at least one

of the polynomials p(t), q(t) is not divisible by t, for each λ �= 0 the pair (B,C + λX)belongs to D2(A) by (1). Therefore (B,C) ∈ D2(A).

However, if r(t) and r′(t) are not divisible by t, then by Lemma 16(a) we can assumethat β � γ. Moreover, the matrix r(A) is invertible and its inverse is a polynomialin A, therefore by Lemma 16(b) the pair (B,C) belongs to D2(A) if and only if thepair (B,C − Aγ−β r′(A)r(A)−1B) belongs to D2(A), and we can assume that r′(t) = 0.Since the matrix C is not a polynomial in A and B, one of the polynomials q′ and p′ isnonzero, therefore the pair (B,C) belongs to D2(A) by the previous paragraph.

Case 2. Assume that k = m. Since for any invertible matrix P ∈ GLn(F) the varietiesN2(A) and N2(P−1AP ) are isomorphic, we can assume that the matrix A is of the block

form A =[ 0 I ··· 0

......

...... I

0

]with k rows and k columns, and each block is a 2×2 matrix. Then

B =[

B1 B2 ··· Bk...

......

... B2B1

]and C =

[C1 C2 ··· Ck

......

...... C2

C1

]for some 2×2 matrices Bi and Ci. Clearly

B1 and C1 have to be nilpotent matrices. By the remark after Lemma 4 of [17] thesematrices B and C can be identified with the polynomials B(t) = B1 +B2t+ · · ·+Bkt

k−1

and C(t) = C1 +C2t+ · · ·+Cktk−1 in M2(F)[t]/tk. We will use this notation in Case 2.

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 249

(1) If B1 or C1 is nonzero matrix, then by Lemma 16(a) we can assume that B1 �= 0.However, then dim(kerA ∩ kerB) = 1, which implies dimF[A,B] = n by Theorem 2 of[13], therefore (B,C) ∈ D2(A).

(2) Assume now that B1 = C1 = 0. If B2 is not a scalar matrix, then the com-mutativity relation of B and C implies that B2C2 = C2B2 which is equivalent toC2 = αI + βB2 for some α, β ∈ F. Moreover, by Lemma 16(b) we can assume thatC2 = 0. Then the commutativity relation implies that B2C3 = C3B2, which againmeans that C3 is linearly dependent of I and B2, and using Lemma 16(b) we can assumethat C3 = 0. Repeating this we can assume that Ci = 0 for each i � k − 1. Since thevariety of all 2 × 2 matrices that are not generic is 3-dimensional, its intersection withany 2-dimensional vector subspace of M2(F) is at least 1-dimensional by Theorem 6 inChapter I, §6.2 of [22]. In particular, there exist μ, ν ∈ F, at least one of them nonzero,such that the matrix μB2 − νCk is not generic. If ν = 0, then B2 is not generic, andsince we can add any multiple of A to B, we can assume that B2 is nilpotent. Thematrix B(t) = B2 + B3t + · · · + Bkt

k−2 is then nilpotent and it commutes with B.Since (B(t), C(t) + λB(t)) ∈ D2(A) for all λ �= 0 by (1), we obtain (B,C) ∈ D2(A).However, if ν �= 0, then by Lemma 16(b) we can subtract μ

νAk−2B from C, therefore

we can assume that Ck is not generic, and by the same lemma we can assume that itis nilpotent. Since C is not a polynomial in A and B, its submatrix Ck is nonzero. Thematrix C(t) = Ck ∈ C(A) is then nilpotent, it commutes with C and by (1) the pair(B(t) + λC(t), C(t)) belongs to D2(A) for each λ �= 0. Therefore (B,C) ∈ D2(A).

It remains to consider the case when B2 is a scalar matrix. Since C is not a polynomialin A, there exists the smallest positive integer l such that Cl is not a scalar matrix.The matrix C(t) = Clt + · · · + Ckt

k−l+1 then commutes with C, and by the previousparagraph (B(t) + λC(t), C(t)) ∈ D2(A) for each λ �= 0. However, then (B,C) ∈ D2(A),which completes the proof of the theorem. �

Corollary 18. If (A,B,C) is any triple of commuting nilpotent n × n matrices withA 2-regular, then (A,B,C) ∈ R(3, n).

Proof. Let A be 2-regular and (X,Y ) ∈ D2(A) arbitrary pair. Then the pair (X, 0)also belongs to D2(A) and Y is a polynomial in A and X by Theorem 1.1 of [16].By Theorem 3.7 of [21] the variety N(2, n) is irreducible for all positive integers n,therefore (A,X) ∈ R(2, n) and (A,X, 0) ∈ R(2, n) × {0}. However, since R(2, n)×{0} ⊆R(3, n), it follows that (A,X, 0) ∈ R(3, n). Now let p ∈ F[t, u] be a polynomial such thatY = p(A,X). The polynomial map ϕ :N(3, n) → N(3, n) defined by ϕ(A1, A2, A3) =(A1, A2, A3 + p(A1, A2)) maps R1(3, n) to itself, therefore ϕ(R(3, n)) ⊆ R(3, n), and inparticular (A,X, Y ) = ϕ(A,X, 0) ∈ R(3, n). We proved that {A} × D2(A) ⊆ R(3, n),therefore {A} ×D2(A) ⊆ R(3, n), and in particular (A,B,C) ∈ R(3, n), since N2(A) isirreducible. �

250 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

Theorem 17 can be viewed as a nilpotent version of Corollary 10 of [17]. However, thesame result for 3-regular matrices, that would be the nilpotent version of Theorem 12 of[25], does not hold as is shown in the following proposition.

Proposition 19. Let A =

⎡⎢⎣0 1 0 0 0 00 0 1 0 0 00 0 0 0 0 00 0 0 0 1 00 0 0 0 0 00 0 0 0 0 0

⎤⎥⎦. Then A is 3-regular nilpotent matrix and the

variety N2(A) is reducible.

Proof. Let V be the variety of all pairs of 6 × 6 matrices⎛⎜⎜⎜⎜⎜⎜⎜⎝

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 a b c d e

0 0 a 0 c 00 0 0 0 0 00 0 f 0 g 00 0 0 0 0 00 0 h 0 i 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦,

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 a′ b′ c′ d′ e′

0 0 a′ 0 c′ 00 0 0 0 0 00 0 f ′ 0 g′ 00 0 0 0 0 00 0 h′ 0 i′ 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦

⎞⎟⎟⎟⎟⎟⎟⎟⎠satisfying cf ′ + eh′ = c′f + e′h and ac′ + cg′ + ei′ = a′c+ c′g + e′i. Clearly V is a propersubvariety of N2(A), and its dimension is at least 16, since it is a subset of F18 definedby 2 equations. However, by Proposition 14 the closure D2(A) has dimension equal todimC(A) − dim kerA + 5 = 16, therefore D2(A) cannot be equal to N2(A), and N2(A)is reducible. �Remark. As we will see in the next section, although the variety N2(A) in the previousproposition is reducible, the variety N(3, 6) is still irreducible. However, to prove this oneneeds to have all three matrices of the triples from {A}×N2(A) perturbed by 1-regularnilpotent commuting matrices, not only two of them.

5. Irreducibility of varieties N(3, n) for n ��� 6

In the previous section we proved that each triple of commuting nilpotent n × n

matrices that contains a 2-regular matrix belongs to R(3, n). In this section we considerall other cases that occur in dimensions 5 and 6. Eventually, we prove that the varietiesN(3, 5) and N(3, 6) are irreducible. We start with two results that hold in any dimension.We prove that a triple of commuting nilpotent n× n matrices belongs to R(3, n) if theJordan canonical forms of the matrices in the triple have only one nonzero Jordan blockor if they have Jordan blocks of orders at most 2. These results will be proved under someadditional assumptions on the perturbability of triples of matrices of higher rank, which,however, have no influence on irreducibility of N(3, n), since in the case of irreducibilityof N(3, n) all triples have to belong to R(3, n). Note that these results are nilpotentversions of Theorems 5.3 and 5.1 of [11].

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 251

Proposition 20. Let k < n and assume that each triple of commuting nilpotent n × n

matrices whose linear span contains a matrix of rank at least k belongs to R(3, n). Let(A,B,C) ∈ N(3, n) be a triple such that the Jordan canonical form of A has one Jordanblock of order k and n− k zero Jordan blocks. Then (A,B,C) ∈ R(3, n).

Proof. If k = n − 1, then the matrix A is 2-regular and the proposition follows fromCorollary 18. Therefore we assume that k � n− 2. By Lemma 7(a) we can assume thatthe matrix A is in the Jordan canonical form. If J denotes the k× k Jordan block, thenB =

[p(J) e1aT

beTk D

]and C =

[p′(J) e1a′ T

b′eTk D′

]for some polynomials with zero constant terms

p, p′ ∈ F[t], some vectors a,a′,b,b′ ∈ Fn−k and some matrices D,D′ ∈ Mn−k(F). By

Lemma 7(c) we can assume that p and p′ are zero polynomials. Moreover, if D or D′ isnonzero, then there exists some linear combination of the matrices A, B and C whichis of rank at least k, and the triple (A,B,C) belongs to R(3, n) by the assumption ofthe proposition. Therefore we can assume that D = D′ = 0. Then the triple (A,B,C)belongs to a subvariety of N(3, n) which is isomorphic to {(a,a′,b,b′) ∈ F

4(n−k); aTb′ =a′ Tb}. This variety is irreducible, since it is defined by one irreducible polynomial.Therefore by Lemma 6 we can assume any open condition on the matrices B and C.We assume that the vectors a and a′ are linearly independent and that aTb �= 0 anda′ Tb′ �= 0. In particular, the vector a is nonzero and by Lemma 7(a) we can assumethat a = e1. Let β = eT1 b and Q =

[ 1 0b I

]∈ GLn−k(F) where b is the vector of

the last n − k − 1 components of 1βb. Then eT1 Q = eT1 and βQe1 = b. Let P =[

I 00 Q

]∈ C(A). By Lemma 7(a) the triple (A,B,C) belongs to R(3, n) if and only if

the triple (A,P−1BP,P−1CP ) does, therefore we can assume that b = βe1. Moreover,by Lemma 7 we can add any multiple of B to C and change the basis of Fn, thereforewe can assume that a′ = e2. The commutativity relation of B and C then implies thateT1 b′ = 0, and similarly as we assumed b = βe1 we can assume also that b′ = −γe2

for some nonzero γ ∈ F. We define (n− k) × (n− k) matrices X =[

β√γ β

√β 0

−γ√β −β

√γ 0

0 0 0

]and

X ′ =[−γ

√β −β

√γ 0

γ√γ γ

√β 0

0 0 0

], where the first two rows and columns are of dimension 1 and the

last ones are of dimension n − k − 2 (possibly 0), and let Y =[ 0 0

0 X

]and Z =

[ 0 00 X′],

where the first rows and columns are of dimension k and the last ones of dimensionn − k. The matrices X and X ′ are of square zero, they commute, and eT1 X ′ = eT2 Xand βX ′e1 = −γXe2, therefore (A,B + λY,C + λZ) ∈ N(3, n) for each λ ∈ F. Sincethe matrix X is nonzero, for each λ �= 0 some linear combination of A and B + λY hasrank k, therefore (A,B + λY,C + λZ) ∈ R(3, n) by the assumption of the proposition.However, then the triple (A,B,C) also belongs to R(3, n). �

Now we deal with triples whose matrices generate a vector space of square zero matri-ces. We first prove a lemma, which can be seen as the nilpotent version of Proposition 4.3of [11]. Our proof is just a slight modification of that of Theorem 1 in [20].

252 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

Lemma 21. Let l be a positive integer, m � 2 and let W ∈ Ml(F) and V ∈ Ml×m(F) bearbitrary. Let

A =

⎡⎣ 0 I 00 0 00 0 0

⎤⎦ and B =

⎡⎣ 0 W V

0 0 00 0 0

⎤⎦

where the first two rows and columns are of dimension l and the last ones are of dimen-

sion m. Then there exists a nonzero nilpotent matrix of the form N =[N1 0 00 N1 00 N2 N3

]which

commutes with A and with B.

Proof. If the matrix V is not injective, then there exists a nonzero matrix X ∈ Mm×l(F)

such that V X = 0. Then the matrix N =[

0 0 00 0 00 X 0

]is nilpotent and it commutes with A

and with B.In the sequel we assume that V is injective, and in particular l � m. The con-

clusion of the lemma clearly remains the same if we conjugate the matrix B by

any invertible matrix of the form P =[P1 0 00 P1 00 P2 P3

](which belongs to the centralizer

of A). In particular, since V is injective, there exist R ∈ Mm×l(F) and Q ∈ GLl(F)such that [ Q−1WQ+Q−1V R Q−1V ] =

[W ′

1 W ′2 0

0 0 I

]for some matrices W ′

1 ∈ Ml−m(F) and

W ′2 ∈ M(l−m)×m(F). Since we can conjugate the matrix B by the matrix P =

[Q 0 00 Q 00 R I

],

we can assume that [ W V ] =[W ′

1 W ′2 0

0 0 I

]where the first row and column are of dimension

l − m and the others are of dimension m. Assume inductively that, for some positive

integer t, [ W V ] =[

W1 W2I

...I

]where the first row and column are of dimension

l − tm and the other t rows and t + 1 columns are of dimension m.Assume first that l − tm �= 0. If W2 is not injective, then (because of m � 2) there

exists a nonzero nilpotent matrix N ′ ∈ Mm(F) such that W2N′ = 0. We define N ′′ =⎡⎣ 0

N ′

N ′...

N ′

⎤⎦ where the first row and column are of dimension l − tm and the other

t rows and columns are of dimension m. The matrix N =[N ′′ 0 00 N ′′ 00 0 N ′

]is then nilpotent

and it commutes with A and with B. On the other hand, if W2 is injective, then l− tm

is not smaller than m and there exist R ∈ Mm×(l−tm)(F) and S ∈ GLl−tm(F) suchthat [ S−1W1S+S−1W2R S−1W2 ] =

[W1 W2 00 0 I

]for some matrices W1 ∈ Ml−(t+1)m(F) and

W2 ∈ M(l−(t+1)m)×m(F). Let W1 =[W1 W2

]∈ Ml−tm(F), W2 =

[ 0 ] ∈ M(l−tm)×m(F),

0 0 I

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 253

Q =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

S

R I

RW1 RW2 I

RW 21 RW1W2 RW2 I

......

.... . . . . .

RW t−11 RW t−2

1 W2 RW t−31 W2 · · · RW2 I

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ Ml(F)

and T = [ RW t1 RW t−1

1 W2 RW t−21 W2 ··· RW2 ] ∈ Mm×l(F). Since we can conjugate the matrix

B by the matrix P =[Q 0 00 Q 00 T I

], we can assume that [ W V ] =

[W1 W2

I...

I

], where the

first row and column are of dimension l − (t + 1)m and the other t + 1 rows and t + 2columns are of dimension m, and we can proceed with the induction.

However, if l− tm = 0, then because of m � 2 there exists a nonzero nilpotent matrix

N ′ ∈ Mm(F). We define N ′′ =[N ′

...N ′

]∈ Ml(F) and N =

[N ′′ 0 00 N ′′ 00 0 N ′

]. Then N is a

nilpotent matrix that commutes with A and with B, which completes the proof of thelemma. �Remark. Note that the above proof would not work for m = 1, even if we assume thatrankB � l − 1, which was assumed in [11] and [20]. As we will see in the proof of thenext proposition, the lemma still holds if m = 1 and rankB � l − 1, but it has to beproved slightly differently.

Proposition 22. Let l be a positive integer, m a nonnegative integer and n = 2l + m.Assume that each triple of commuting nilpotent n×n matrices whose linear span containseither a matrix of rank at least l + 1 or a matrix of rank l with nonzero square belongsto R(3, n). Let (A,B,C) ∈ N(3, n) be a triple such that the Jordan canonical form of Ahas l Jordan blocks of order 2 and m zero Jordan blocks. Then (A,B,C) ∈ R(3, n).

Proof. By Lemma 7(a) we can conjugate the matrices by any invertible n × n matrix,

therefore we can assume that A =[

0 I 00 0 00 0 0

], where the first two block rows and columns

are of dimension l and the last ones are of dimension m (possibly 0). Then B =[D E F0 D 00 G H

]and C =

[D′ E′ F ′

0 D′ 00 G′ H′

]for some matrices D,D′, E,E′ ∈ Ml(F), F, F ′ ∈ Ml×m(F), G,G′ ∈

Mm×l(F) and H,H ′ ∈ Mm(F). By the assumption of the proposition we can assume thateach linear combination of the matrices A, B and C is a square-zero matrix of rank atmost l. The second condition immediately implies that H = H ′ = 0 if m > 0.

If D �= 0 or D′ �= 0, then by Lemma 7(b) we can assume that D �= 0 and thateither the matrices D and D′ are linearly independent or D′ = 0. The coefficient at

λ in the polynomial (B + λA)2 is equal to[

0 2D 00 0 0

]. If charF �= 2, then there exists

0 0 0

254 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

λ ∈ F such that (B + λA)2 �= 0, and therefore (A,B,C) ∈ R(3, n) by the assumptionof the proposition. Therefore we can assume that charF = 2. The condition B2 = 0is equivalent to the equations D2 = 0, DE + ED + FG = 0, DF = 0 and GD = 0.In particular, D and D′ are square-zero matrices, and the condition that either D andD′ are linearly independent or D′ = 0 implies that the algebra F[D] = F · I + F · Dis not contained in the algebra F[D′] = F · I + F · D′. Recall that for any matrix X ∈Ml(F) we denoted by C(X) the centralizer of X. Similarly we denote by C(C(X)) thecentralizer of the algebra C(X) in Ml(F) (i.e. the set of all matrices from Ml(F) thatcommute with all elements of C(X)) and by Z(C(X)) the center of the algebra C(X) (i.e.Z(C(X)) = C(C(X)) ∩ C(X)). Theorem 7 in Chapter 1 of [23] yields Z(C(X)) = F[X]for all X ∈ Ml(F), and since each matrix that commutes with all elements of C(X)clearly belongs to C(X), we obtain C(C(X)) = F[X] for all X ∈ Ml(F). Since thealgebra F[D] is not contained in F[D′], the centralizer C(D′) is not contained in C(D).Therefore there exists a matrix Y ∈ C(D′)\C(D). Note that then DY + Y D �= 0, since

charF = 2. The matrix X =[

0 Y 00 0 00 0 0

]then commutes with A and C, and (B + λX)2 �= 0

for λ �= 0. The assumption of the proposition then implies that (A,B+λX,C) ∈ R(3, n)for all λ �= 0, hence (A,B,C) ∈ R(3, n). In the rest of the proof we can therefore assumethat D = D′ = 0.

Assume first that m = 0. Since the variety of all l× l matrices that are not generic is(l2−1)-dimensional, its intersection with each 2-dimensional vector subspace of Ml(F) isat least 1-dimensional by Theorem 6 in Chapter I, §6.2 of [22]. In particular, there existλ, μ ∈ F, not both of them zero, such that the matrix λE+μE′ is not generic. Moreover,by Lemma 7(b) we can assume that E is not generic. Therefore it commutes with somenonzero nilpotent matrix N ′ ∈ Ml(F). The matrix N =

[N ′ 00 N ′

]is then nilpotent and it

commutes with B, so (A,B,C + λN) ∈ N(3, n) for each λ ∈ F. Since N ′ �= 0, the triple(A,B,C + λN) belongs to R(3, n) for all λ �= 0 by the previous paragraph, therefore(A,B,C) ∈ R(3, n).

In the rest of the proof we will assume that m > 0. Since B2 = 0, it follows thatFG = 0. Moreover, for each λ ∈ F the rank of the matrix λA+B is at most l, which im-plies that rank

[λI+E F

G 0

]� l. In particular, for each i, j = 1, . . . ,m, det

[λI+E Fei

eTj G 0

]= 0.

The coefficient at λl−1 in this determinant is equal to

l∑k=1

eTj GekeTk Fei = eTj GFei,

therefore eTj GFei = 0 for all i, j = 1, . . . ,m, i.e. GF = 0.Assume first that there exist μ, ν ∈ F such that the matrices μG+νG′ and μF+νF ′ are

both nonzero. By Lemma 7(b) we can then assume that F and G are nonzero matrices.Since GF = 0 and FG = 0, it follows that m > 1 and that the rank of the matrices F

and G is strictly smaller than min{m, l}. By Lemma 7(a) we can conjugate the matricesB and C by any invertible matrix in the centralizer of A, therefore we can assume that

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 255

F =∑k

i=1 eieTi for some k � min{m, l}− 1. In particular, the condition FG = 0 implies

that eT1 G = 0. The matrix N =[ 0 0 0

0 0 00 0 emeT

1

]is then nilpotent and it commutes with A

and with B. Moreover, for each λ �= 0 there exists a linear combination of A and C+λN

which has rank at least l + 1, so (A,B,C + λN) ∈ R(3, n) by the assumption of theproposition, and therefore (A,B,C) ∈ R(3, n).

We can now assume that for all μ, ν ∈ F we have μF + νF ′ = 0 or μG + νG′ = 0,which is equivalent to F = F ′ = 0 or G = G′ = 0, and by Lemma 7(e) we can assumethat G = G′ = 0. The triple (A,B,C) belongs to the 2l(m + l)-dimensional affine spaceof all triples of n× n matrices such that the first matrix of the triple is equal to A andthe other two matrices have nonzero entries only in the last two block columns of thefirst block row. This is an irreducible subvariety of N(3, n), therefore by Lemma 6 we canassume any open condition on A, B and C, and we assume that some linear combinationof F and F ′ is of full rank (i.e. equal to min{m, l}).

If m � 2, then we can use the previous lemma to obtain a nonzero nilpotent n × n

matrix of the form N =[N1 0 00 N1 00 N2 N3

]that commutes with A and B, therefore (A,B,C +

λN) ∈ N(3, n) for all λ ∈ F. Since N is nonzero and some linear combination of F andF ′ has full rank, the triple (A,B,C + λN) belongs to R(3, n) for all λ �= 0 by the casesalready considered in this proof. Therefore the triple (A,B,C) also belongs to R(3, n).

However, if m = 1, then by Corollary 2.1 of [27] there exists some nonzero linearcombination of A, B and C that has rank at most l − 1, and by Lemma 7(b) we canassume that rankB � l−1. Therefore there exists some nonzero vector x ∈ F

l such thatxTE = 0 and xTF = 0. Moreover, since [E F ] is l×(l+1) matrix of rank at most l−1,its kernel is at least 2-dimensional. In particular, there exist y ∈ F

l and ζ ∈ F, not both

of them zero, such that xTy = 0 and Ey + ζF = 0. The matrix N =[

yxT 0 00 yxT 00 ζxT 0

]is

then nilpotent and it commutes with A and B. As in the previous paragraph the triple(A,B,C + λN) belongs to R(3, n) for each λ �= 0, therefore (A,B,C) ∈ R(3, n), whichcompletes the proof of the proposition. �

To prove irreducibility of the variety N(3, 6) only the case of Jordan blocks of orders3, 2 and 1 is left to be considered. We will prove that triples of commuting nilpotentmatrices with such Jordan structure also belong to R(3, 6). However, if the matrix A

is such a matrix, then the variety N2(A) is reducible, cf. Proposition 19. This meansthat we have to perturb all three matrices in the triple, not only two of them, as itwas done in all previous cases. The perturbed matrices can be computed explicitly, butexpressing them in terms of the original matrices is complicated. Therefore we prefer togive a geometric proof of their existence.

Proposition 23. Let (A,B,C) be a triple of commuting nilpotent 6 × 6 matrices. If theJordan canonical form of A has one Jordan block of order 3, one Jordan block of order2 and one zero Jordan block, then the triple (A,B,C) belongs to R(3, 6).

256 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

Proof. By Lemma 7(a) we can assume that the matrix A is in the Jordan canonicalform. Since the matrices B and C are nilpotent and they commute with A, they look

like B =

⎡⎢⎢⎣0 a b c d e0 0 a 0 c 00 0 0 0 0 00 f g 0 h i0 0 f 0 0 00 0 j 0 k 0

⎤⎥⎥⎦ and C =

⎡⎢⎢⎣0 a′ b′ c′ d′ e′

0 0 a′ 0 c′ 00 0 0 0 0 00 f ′ g′ 0 h′ i′

0 0 f ′ 0 0 00 0 j′ 0 k′ 0

⎤⎥⎥⎦. If there is some λ ∈ F such that

c + λc′ and f + λf ′ are both nonzero, then the matrix B + λC has rank at least 4and (A,B,C) ∈ R(3, 6) by Corollary 18. Therefore we can in the sequel assume thatc = c′ = 0 or f = f ′ = 0, and by Lemma 7(e) we can assume that f = f ′ = 0. We willconsider two cases.

Case 1. Assume that c �= 0 or c′ �= 0. By Lemma 7(b) we can assume that c �= 0. Ifi + λi′ �= 0 for some λ ∈ F, then we can by Lemma 7(b) assume that i �= 0. However,then dim(kerA ∩ kerB) = 1, so Theorem 2 of [13] implies that dimF[A,B] = 6, i.e.(B,C) ∈ D2(A), and therefore (A,B,C) ∈ R(3, 6) (see the proof of Corollary 18). Inthe rest of Case 1 we can therefore assume that i = i′ = 0. To simplify the notation of

the matrices we will change the basis of F6. Let P =

⎡⎢⎣1 0 0 0 0 00 1 0 0 0 00 0 0 0 1 00 0 1 0 0 00 0 0 0 0 10 0 0 1 0 0

⎤⎥⎦. By Lemma 7(a) the

triple (A,B,C) belongs to R(3, 6) if and only if the triple (P−1AP,P−1BP,P−1CP )does, therefore we will assume that the matrices A, B and C are in the following blockform:

A =

⎡⎣ 0 eT1 00 0 I

0 0 0

⎤⎦ , B =

⎡⎣ 0 bT b′ T

0 0 B

0 0 0

⎤⎦ and C =

⎡⎣ 0 cT c′ T0 0 C

0 0 0

⎤⎦where the first rows and columns are of dimension 1, the second ones of dimension 3 andthe last ones of dimension 2, and

I =

⎡⎣ 1 00 10 0

⎤⎦ , B =

⎡⎣ a c

g h

j k

⎤⎦ , C =

⎡⎣ a′ c′

g′ h′

j′ k′

⎤⎦ ,b =

⎡⎣ ace

⎤⎦ , c =

⎡⎣ a′c′e′

⎤⎦ , b′ =[b

d

], c′ =

[b′

d′

].

The commutativity relation of B and C is equivalent to bT C = cT B. Since bTe2 �= 0, theentries of the second row of C can be expressed as rational functions in b, c, B and theother two rows of C, therefore the triple (A,B,C) belongs to a rationally parameterizedsubvariety of N(3, 6), which is irreducible by Proposition 6 in Chapter 4, §5 of [4]. Inparticular, by Lemma 6 we can assume any open condition on the matrices A, B and C.We assume that the vectors e1, b and c are linearly independent. Therefore there exists a

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 257

matrix Q ∈ M3×2(F) such that eT1 Q = 0, bTQ = b′ T and cTQ = c′ T . Let S =[

1 0 00 I Q0 0 I

].

Since by Lemma 7(a) the triple (A,B,C) belongs to R(3, 6) if and only if the triple(SAS−1, SBS−1, SCS−1) does, we can assume that b′ = c′ = 0.

Let V be the variety of all triples (X1, X2, X3) ∈ N(3, 6) such that for each i = 1, 2, 3the matrix Xi is of the form

Xi =

⎡⎣ 0 xTi 0

0 Yi Zi

0 0 ζiJ

⎤⎦ (1)

for some xi ∈ F3, Yi ∈ M3(F), Zi ∈ M3×2(F) and ζi ∈ F, where J denotes the nilpotent

2 × 2 Jordan block. Furthermore, let U be the open subset of V consisting of all triples(X1, X2, X3) ∈ V with X1 1-regular and x1,x2,x3 linearly independent, let U1 be the setof all 6 × 6 matrices X1 such that (X1, X2, X3) ∈ U for some X2, X3 ∈ M6(F), and letU ′

1 be the set of all 1-regular nilpotent matrices of the form (1). Clearly U1 ⊆ U ′1 and it

is also clear that the matrix Xi of the form (1) is nilpotent if and only if Yi is nilpotent.

Moreover, since X51 =

[0 0 ζ1xT

1 Y 21 Z1J

0 0 00 0 0

], 1-regularity of X1 implies ζ1xT

1 Y21 Z1e1 �= 0,

therefore we can define rational map ϕ :U ′1 × F

6 → V by

ϕ(X,α, β, γ, α′, β′, γ′)=(X,αX + βX2 + γX3 − βxTZe1 + γxTY Ze1

xTY 2Ze1X4

+(

(βxTZe1 + γxTY Ze1)(ζxTY Ze1 + xTY 2Ze2)ζ(xTY 2Ze1)2

− βxTZe2 + γ(ζxTZe1 + xTY Ze2)ζxTY 2Ze1

)X5, α′X + β′X2 + γ′X3

− β′xTZe1 + γ′xTY Ze1

xTY 2Ze1X4 +

((β′xTZe1 + γ′xTY Ze1)(ζxTY Ze1 + xTY 2Ze2)

ζ(xTY 2Ze1)2

− β′xTZe2 + γ′(ζxTZe1 + xTY Ze2)ζxTY 2Ze1

)X5)

where X =[

0 xT 00 Y Z0 0 ζJ

]. This map is clearly injective.

If (X1, X2, X3) ∈ U , then X2 and X3 are polynomials in 1-regular matrix X1,i.e. (X1, X2, X3) = ϕ(X1, α, β, γ, α

′, β′, γ′) for some α, β, γ, α′, β′, γ′ ∈ F. Since x1,x2 and x3 are linearly independent, we obtain that xT

1 , xT1 Y1 and xT

1 Y21 are lin-

early independent and that (β, γ) and (β′, γ′) are linearly independent. Conversely, ifa nilpotent matrix X1 of the form (1) is 1-regular, xT

1 ,xT1 Y1,xT

1 Y21 are linearly inde-

pendent, (β, γ), (β′, γ′) ∈ F2 are linearly independent and α, α′ ∈ F are arbitrary, then

ϕ(X1, α, β, γ, α′, β′, γ′) ∈ U . Therefore U1 consists of all 1-regular nilpotent matrices of

258 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

the form (1) such that xT1 , xT

1 Y1 and xT1 Y

21 are linearly independent, and U = ϕ(U1×U2)

where U2 = {(α, β, γ, α′, β′, γ′) ∈ F6; βγ′ �= β′γ}. In particular, U2 is open in F

6 andU1 is open in the variety of all nilpotent matrices of the form (1) which is isomorphic toF

10 ×N3. Moreover, U1 is nonempty, since it contains the nilpotent 6 × 6 Jordan block.F

10 ×N3 is irreducible, therefore U1 is irreducible and dimU1 = 10 + dimN3 = 16.Since the variety U1 ×F

6 is irreducible, Proposition 14 of [26] implies that the varietyU = ϕ(U1 × U2) is also irreducible. Moreover, since ϕ is injective, Theorem 11.12 of [10]implies that dimU = 6 + dimU1 = 22.

Now we define

V ′ ={(x1, Z1,x2, Z2,x3, Z3) ∈

(F

3 ×M3×2(F))3;

xTi Zj = xT

j Zi for all distinct i and j}.

Note that this variety is isomorphic to the subvariety of V consisting of all triples of theform ⎛⎝⎡⎣ 0 xT

1 00 0 Z10 0 0

⎤⎦ ,⎡⎣ 0 xT

2 00 0 Z20 0 0

⎤⎦ ,⎡⎣ 0 xT

3 00 0 Z30 0 0

⎤⎦⎞⎠for some xi ∈ F

3 and some Zi ∈ M3×2(F). Furthermore, let U ′ be the subset of V ′

consisting of such 6-tuples in V ′ that x1, x2 and x3 are linearly independent, and let

W ={(Z1, Z2, Z3) ∈ M3×2(F)3; eTi Zj = eTj Zi for all distinct i and j

}.

The variety W is a vector space, therefore it is irreducible and of dimension 12. Moreover,if (x1, Z1,x2, Z2,x3, Z3) ∈ U ′, then there exists a unique matrix S ∈ GL3(F) suchthat eTi S = xT

i for i = 1, 2, 3. The rational map ψ :GL3(F) × W → U ′ defined byψ(S,Z1, Z2, Z3) = (STe1, S

−1Z1, STe2, S

−1Z2, STe3, S

−1Z3) is therefore a birationalequivalence, so U ′ is irreducible by Proposition 6 in Chapter 4, §5 of [4], and dimU ′ =dimW + dimGL3(F) = 21 by Corollary 7 in Chapter 9, §5 of [4].

Now we define the projection π :U → U ′ by

π

⎛⎝⎡⎣ 0 xT1 0

0 Y1 Z10 0 ζ1J

⎤⎦ ,⎡⎣ 0 xT

2 00 Y2 Z20 0 ζ2J

⎤⎦ ,⎡⎣ 0 xT

3 00 Y3 Z30 0 ζ3J

⎤⎦⎞⎠ = (x1, Z1,x2, Z2,x3, Z3).

Assume that the projection π is not dominant (i.e. the image π(U) is not dense inU ′). Then π(U) is a proper subvariety of U ′, and the irreducibility of U ′ implies thatdim π(U) � dimU ′ − 1 = 20. Theorem 11.12 of [10] then implies that any component ofany fibre of the projection π has dimension at least dimU − dim π(U) � 2.

However, let xi = ei for i = 1, 2, 3,

Z1 =

⎡⎣ 0 10 0

⎤⎦ , Z2 =

⎡⎣ 0 01 0

⎤⎦ and Z3 =

⎡⎣ 1 00 1

⎤⎦ ,

1 0 0 1 0 0

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 259

and assume that π(X1, X2, X3) = (x1, Z1,x2, Z2,x3, Z3) where for each i = 1, 2, 3, Xi =[0 xT

i 00 Yi Zi

0 0 ζiJ

]for some Yi ∈ M3(F) and ζi ∈ F, and at least one of Y1, Y2, Y3, ζ1, ζ2, ζ3 is

nonzero. Then

xTi Yj = xT

j Yi, (2)

YiZj + ζjZiJ = YjZi + ζiZjJ, (3)

YiYj = YjYi (4)

for all distinct i and j. Moreover, the matrices Yi are nilpotent and in particular

Tr(Yi) = 0 (5)

for each i = 1, 2, 3. Solving linear equations (2), (3) and (5) we obtain

Y1 =

⎡⎣ α β γ

γ −2α β

β − δ γ α

⎤⎦ , Y2 =

⎡⎣ γ −2α β

β − δ −2γ −2α−2α β − δ γ

⎤⎦ ,Y3 =

⎡⎣β − δ γ α

−2α β − δ γ

4γ −2α β − δ

⎤⎦ ,ζ1 = δ, ζ2 = −3α, ζ3 = 3γ and 3(β − δ) = 0. Now we have to consider different cases forthe characteristic of the field F.

If charF �= 2, 3, then first we get δ = β. A short calculation shows that (4) is equivalentto the equations 2αβ = 3γ2 and 2βγ = −3α2. Since at least one of α, β and γ is nonzero,it follows that β �= 0 and

Y1 =

⎡⎣ βω β −32βω

2

−32βω

2 −2βω β

0 −32βω

2 βω

⎤⎦ , Y2 =

⎡⎣−32βω

2 −2βω β

0 3βω2 −2βω−2βω 0 −3

2βω2

⎤⎦ ,Y3 =

⎡⎣ 0 −32βω

2 βω

−2βω 0 −32βω

2

−6βω2 −2βω 0

⎤⎦ ,ζ1 = β, ζ2 = −3βω and ζ3 = −9

2βω2, where ω(27ω3 − 8) = 0. It can be verified

that the matrices Y1, Y2 and Y3 are indeed nilpotent and X1 is 1-regular for β �= 0.The fibre π−1(x1, Z1,x2, Z2,x3, Z3) is therefore nonempty and it is equal to a unionof a finite number of 1-dimensional affine spaces, hence it is 1-dimensional, which is acontradiction.

If charF = 2, then we again obtain δ = β. The matrix Y2 is then upper triangularwith the diagonal (γ, 0, γ), and since it is nilpotent, it follows that γ = 0. However,

260 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

then the matrix Y1 is upper triangular with the diagonal (α, 0, α), therefore α = 0.

Hence Y1 =[

0 β 00 0 β0 0 0

], Y2 =

[0 0 β0 0 00 0 0

], Y3 = 0, ζ1 = β and ζ2 = ζ3 = 0, i.e. the fibre

π−1(x1, Z1,x2, Z2,x3, Z3) is 1-dimensional affine space, which is again a contradiction.However, if charF = 3, then the relation (4) implies that αδ = γδ = 0. Assume

that (X1, X2, X3) ∈ U . Then δ �= 0, so α = γ = 0. However, then Y3 = (β − δ)I,so δ = β and we get the same solution as in the case of charF = 2. There-fore dim(U ∩ π−1(x1, Z1,x2, Z2,x3, Z3)) = 1. However, this set is open in the fibreπ−1(x1, Z1,x2, Z2,x3, Z3), therefore its closure is a union of some irreducible compo-nents of π−1(x1, Z1,x2, Z2,x3, Z3). Hence, there exists a component of this fibre whichis 1-dimensional, which is again a contradiction.

In each characteristic we obtained a contradiction, therefore the projection π is dom-inant. Then the continuity of the projection π implies that U ′ = π(U) ⊆ π(U), i.e.π(U) is also dense in U ′. If (x1, Z1,x2, Z2,x3, Z3) ∈ π(U) is arbitrary, then there existY1, Y2, Y3 ∈ N3 and ζ1, ζ2, ζ3 ∈ F such that⎛⎝⎡⎣ 0 xT

1 00 Y1 Z10 0 ζ1J

⎤⎦ ,⎡⎣ 0 xT

2 00 Y2 Z20 0 ζ2J

⎤⎦ ,⎡⎣ 0 xT

3 00 Y3 Z30 0 ζ3J

⎤⎦⎞⎠ ∈ U ⊆ R1(3, 6).

Moreover, since the line{[

0 xT1 0

0 λY1 Z10 0 λζ1J

]; λ ∈ F

}intersects the open set of all 1-regular

matrices, all except finitely many matrices on this line are 1-regular, i.e.⎛⎝⎡⎣ 0 xT1 0

0 λY1 Z10 0 λζ1J

⎤⎦ ,⎡⎣ 0 xT

2 00 λY2 Z20 0 λζ2J

⎤⎦ ,⎡⎣ 0 xT

3 00 λY3 Z30 0 λζ3J

⎤⎦⎞⎠ ∈ R1(3, 6)

for all except finitely many scalars λ ∈ F, and therefore⎛⎝⎡⎣ 0 xT1 0

0 0 Z10 0 0

⎤⎦ ,⎡⎣ 0 xT

2 00 0 Z20 0 0

⎤⎦ ,⎡⎣ 0 xT

3 00 0 Z30 0 0

⎤⎦⎞⎠ ∈ R(3, 6).

We proved that the set⎧⎨⎩⎛⎝⎡⎣ 0 xT

1 00 0 Z10 0 0

⎤⎦ ,⎡⎣ 0 xT

2 00 0 Z20 0 0

⎤⎦ ,⎡⎣ 0 xT

3 00 0 Z30 0 0

⎤⎦⎞⎠ ; (x1, Z1,x2, Z2,x3, Z3) ∈ π(U)

⎫⎬⎭is a subset of R(3, 6), and since π(U) = U ′, the set⎧⎨⎩⎛⎝⎡⎣ 0 xT

1 00 0 Z10 0 0

⎤⎦ ,⎡⎣ 0 xT

2 00 0 Z20 0 0

⎤⎦ ,⎡⎣ 0 xT

3 00 0 Z30 0 0

⎤⎦⎞⎠ ; (x1, Z1,x2, Z2,x3, Z3) ∈ U ′

⎫⎬⎭is also a subset of R(3, 6). In particular (A,B,C) ∈ R(3, 6), which proves Case 1.

N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262 261

Case 2. Assume that c = c′ = 0. If there exists λ ∈ F such that i+λi′ and k+λk′ areboth nonzero, then there exists μ ∈ F such that rank(μA+B+λC) = 4, and (A,B,C) ∈R(3, 6) by Corollary 18. Therefore we can assume that i = i′ = 0 or k = k′ = 0, andby Lemma 7(e) we can assume that k = k′ = 0. The commutativity relation of B andC is then equivalent to the equations ej′ = e′j and ij′ = i′j. If j �= 0 or j′ �= 0, thenby Lemma 7(b) we can assume that j = 1 and j′ = 0, and the commutativity relationof B and C implies that e′ = i′ = 0. The triple (A,B,C) then belongs to some affinespace, therefore by Lemma 6 we can assume that i �= 0. The matrix X = e6eT5 commuteswith A and with C, therefore (A,B + λX,C) ∈ N(3, 6) for each λ ∈ F. Moreover, foreach λ �= 0 there exists μ ∈ F such that rank(B + λX + μA) = 4, therefore the triple(A,B + λX,C) belongs to R(3, 6) by Corollary 18. Hence (A,B,C) ∈ R(3, 6).

It remains to consider the case when j = j′ = 0. By Lemma 7(b) we can exchange B

and C or add any multiple of B to C, therefore we can assume that i′ = 0. The triple(A,B,C) then belongs to a 13-dimensional affine space which is an irreducible subvarietyof N(3, 6), and by Lemma 6 we can assume any open condition on B and C. We will

assume that e′ �= 0. Then the matrix X =

⎡⎢⎢⎣0 0 0 e′ 0 00 0 0 0 e′ 00 0 0 0 0 00 0 0 0 0 00 0 0 0 0 00 0 g′ 0 h′ 0

⎤⎥⎥⎦ commutes with A and with C,

and (A,B + λX,C) ∈ R(3, 6) for each λ �= 0 by Case 1. This implies that the triple(A,B,C) also belongs to R(3, 6), which completes the proof of the proposition. �

Now we can prove the main theorem of this section: irreducibility of N(3, n) for n � 6.Note that the new cases are n = 5 and n = 6, since irreducibility of N(3, n) for n � 3was proved in [18] and irreducibility of N(3, 4) was proved in [28].

Theorem 24. The varieties N(3, n) are irreducible for n � 6.

Proof. We have to prove that each triple from N(3, n) belongs to R(3, n). Let (A,B,C) ∈N(3, n) be any triple. By Lemma 7(b) we can assume that for each i � n − 1 the rankof Ai is not smaller than the rank of the i-th power of any linear combination of A, Band C. If A is 2-regular, then (A,B,C) ∈ R(3, n) by Corollary 18. Otherwise the Jordancanonical form of A either has only one nonzero Jordan block or all Jordan blocks oforder at most 2 or Jordan blocks of sizes 3, 2 and 1. In all cases, the tripe (A,B,C)belongs to R(3, n) by respectively Propositions 20, 22 and 23. �References

[1] V. Baranovsky, The variety of pairs of commuting nilpotent matrices is irreducible, Transform.Groups 6 (2001) 3–8.

[2] J. Barria, P.R. Halmos, Vector bases for two commuting matrices, Linear Multilinear Algebra 27(1990) 147–157.

[3] R. Basili, On the irreducibility of commuting varieties of nilpotent matrices, J. Algebra 268 (2003)58–80.

262 N.V. Ngo, K. Šivic / Linear Algebra and its Applications 452 (2014) 237–262

[4] D. Cox, J. Little, D. O’Shea, Ideals, Varieties, and Algorithms, 2nd edition, Springer-Verlag, NewYork, 1997.

[5] M. Gerstenhaber, On dominance and varieties of commuting matrices, Ann. of Math. 73 (1961)324–348.

[6] R.M. Guralnick, A note on commuting pairs of matrices, Linear Multilinear Algebra 31 (1992)71–75.

[7] R.M. Guralnick, N.V. Ngo, Reducibility of nilpotent commuting varieties, preprint, available athttp://arxiv.org/pdf/1308.2420.pdf, 2013.

[8] R.M. Guralnick, B.A. Sethuraman, Commuting pairs and triples of matrices and related varieties,Linear Algebra Appl. 310 (2000) 139–148.

[9] Y. Han, Commuting triples of matrices, Electron. J. Linear Algebra 13 (2005) 274–343.[10] J. Harris, Algebraic Geometry, Grad. Texts in Math., Springer-Verlag, New York, 1992.[11] J. Holbrook, M. Omladič, Approximating commuting operators, Linear Algebra Appl. 327 (2001)

131–149.[12] A.A. Kirillov, Y.A. Neretin, The variety An of n-dimensional Lie algebra structures, Amer. Math.

Soc. Transl. Ser. 2 137 (1987) 21–30, fourteen papers translated from Russian.[13] T. Košir, On the structure of commutative matrices II, Linear Algebra Appl. 261 (1997) 293–305.[14] T.J. Laffey, S. Lazarus, Two-generated commutative matrix subalgebras, Linear Algebra Appl. 147

(1991) 249–273.[15] T.S. Motzkin, O. Taussky, Pairs of matrices with property L II, Trans. Amer. Math. Soc. 80 (1955)

387–401.[16] M.G. Neubauer, D.J. Saltman, Two-generated commutative subalgebras of Mn(F ), J. Algebra 164

(1994) 545–562.[17] M.G. Neubauer, B.A. Sethuraman, Commuting pairs in the centralizers of 2-regular matrices, J. Al-

gebra 214 (1999) 174–181.[18] N.V. Ngo, Commuting varieties of r-tuples over Lie algebra, J. Pure Appl. Algebra 218 (2014)

1400–1417.[19] K.C. O’Meara, J. Clark, C.I. Vinsonhaler, Advanced Topics in Linear Algebra: Weaving Matrix

Problems Through the Weyr Form, Oxford University Press, New York, 2011.[20] M. Omladič, A variety of commuting triples, Linear Algebra Appl. 383 (2004) 233–245.[21] A. Premet, Nilpotent commuting varieties of reductive Lie algebras, Invent. Math. 154 (2003)

653–683.[22] I.R. Shafarevich, Basic Algebraic Geometry 1, 2nd edition, Springer-Verlag, Berlin, 1994.[23] D.A. Suprunenko, R.I. Tyshkevich, Commutative Matrices, Academic Paperbacks, Academic Press,

New York, 1968.[24] K. Šivic, On varieties of commuting triples, Linear Algebra Appl. 428 (2008) 2006–2029.[25] K. Šivic, On varieties of commuting triples II, Linear Algebra Appl. 437 (2012) 461–489.[26] K. Šivic, On varieties of commuting triples III, Linear Algebra Appl. 437 (2012) 393–460.[27] R. Westwick, Spaces of linear transformations of equal rank, Linear Algebra Appl. 5 (1972) 49–64.[28] H.-W. Young, Components of algebraic sets of commuting and nearly commuting matrices, PhD

thesis, University of Michigan, 2010.