This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On Linear spaces of of matrices bounded rank

Hang Huang and J. M. Landsberg Department of Mathematics, Texas A&M University, College Station, TX 77843-3368, USA [email protected]
Abstract.

Motivated by questions in theoretical computer science and quantum information theory, we study the classical problem of determining linear spaces of matrices of bounded rank. Spaces of bounded rank three were classified in 1983, and it has been a longstanding problem to classify spaces of bounded rank four. Before our study, no non-classical example of such a space was known. We exhibit two non-classical examples of such spaces and give the full classification of basic spaces of bounded rank four. There are exactly four such up to isomorphism. We also take steps to bring together the methods of the linear algebra community and the algebraic geometry community used to study spaces of bounded rank.

2010 Mathematics Subject Classification:
68Q17; 14L30, 15A69, 15A30
Huang supported by NSF grant DMS2302375, Landsberg supported by NSF grant AF-2203618

1. Introduction

A linear subspace E𝐛𝕔E\subset\mathbb{C}^{\mathbf{b}}{\mathord{\otimes}}\mathbb{C}^{\mathbb{c}} is of bounded rank if for all eEe\in E, rank(e)<min{𝐛,𝕔}{\mathrm{rank}}(e)<\operatorname{min}\{\mathbf{b},{\mathbb{c}}\}. We say a space of bounded rank has bounded rank rr if rr is the maximal rank of an element of EE. Fix 𝕒=dimE{\mathbb{a}}=\operatorname{dim}E.

Example 1.1.

Let EE be of the form (0)\begin{pmatrix}*&*\\ *&0\end{pmatrix} where the blocking is (k1,𝐛k1)×(k2,𝕔k2)(k_{1},\mathbf{b}-k_{1})\times(k_{2},{\mathbb{c}}-k_{2}). Then EE has bounded rank (at most) k1+k2k_{1}+k_{2}. These are called compression spaces, or more precisely (k1,k2)(k_{1},k_{2})-compression spaces.

Example 1.2.

Let E=Λ22p+12p+12p+1E=\Lambda^{2}\mathbb{C}^{2p+1}\subset\mathbb{C}^{2p+1}{\mathord{\otimes}}\mathbb{C}^{2p+1} be the space of 2p+1×2p+12p+1\times 2p+1 skew symmetric matrices. Then EE is of bounded rank 2p2p as the rank of a skew-symmetric matrix is always even.

Example 1.3.

Let EHom(E,Λ2E)E\subset\operatorname{Hom}(E,\Lambda^{2}E) be given by e{vev}e\mapsto\{v\mapsto e\wedge v\}, which has bounded rank 𝕒1{\mathbb{a}}-1, the kernel of the map associated to ee is the line through ee.

We will refer to these examples as the classical spaces of bounded rank.

The study of spaces of bounded rank dates back at least to Flanders in 1962 [13], who solved a conjecture on the maximal dimension of such a space posed by Marcus. Once the problem is stated, the case 𝕒=2{\mathbb{a}}=2 follows immediately from the Kronecker-Weierstrass normal form for pencils of matrices.

Atkinson-Lloyd [3] introduced the notion of primitive spaces and proposed the classification of such by rank. As was known for a long time, the case r=1r=1 consists of the (0,1)(0,1) and (1,0)(1,0) compression spaces. They classified the r=2r=2 case, where the only primitive space is Example 1.2 with p=1p=1. In [2] Atkinson carried out the classification for r=3r=3, the only primitive examples are Example 1.3 and its projections. In particular, there are no non-classical examples of spaces of bounded rank when r3r\leq 3. In the same paper he observed that if one allows rr to be large, there are “many”  such spaces. This work is reviewed in §4.1.

Sylvester [26] introduced language from geometry (vector bundles, Chern classes) to study the more restrictive problem of spaces of constant rank rr.

Eisenbud-Harris [12] independently introduced these tools, where for the general question one must deal with sheaves rather than vector bundles. They proposed a refinement of the classification problem to the study of basic spaces (which in particular, disallows projections of primitive spaces) and stated that they were unaware of a non-classical example of a basic space of bounded rank four. They also refined the observation that there are many such spaces for rr large by observing that such spaces arise as matrices appearing in linear parts of the minimal free resolutions of sufficiently general projective curves. In particular, there are nontrivial moduli of basic spaces for rr large. Their work is reviewed in §4.2.

To our knowledge, the first example of a non-classical space of bounded rank is due to Westwick in 1996 [28], which is even of constant rank (namely constant rank 88). Since then numerous explicit examples have been found. See [19] and [23] for two recent contributions in the special case of constant rank.

Acknowledgements

We thank Austin Conner, Harm Derksen, David Eisenbud, Mark Green, Joe Harris, Laurent Manivel, Mihnea Popa, Jerzy Weyman, and Derek Wu for useful conversations. Landsberg especially thanks Harris for inviting him to Harvard where this project began to take shape.

2. Results

Basic spaces are reviewed in §4.2. All spaces of bounded rank may be deduced from the basic spaces.

Theorem 2.1.

[Main Theorem] Up to isomorphism, there are exactly four basic spaces of bounded rank four:

  1. (I)

    EΛ2555E\cong\Lambda^{2}\mathbb{C}^{5}\subset\mathbb{C}^{5}{\mathord{\otimes}}\mathbb{C}^{5},

  2. (II)

    E5Hom(5,Λ25)E\cong\mathbb{C}^{5}\subset\operatorname{Hom}(\mathbb{C}^{5},\Lambda^{2}\mathbb{C}^{5}),

  3. (III)
    E=(a1a3a5a1a4a6a1a20a10a2a20a3a5000a2a4a600)66,E=\begin{pmatrix}a_{1}&&&&-a_{3}&-a_{5}\\ &a_{1}&&&-a_{4}&-a_{6}\\ &&a_{1}&&a_{2}&0\\ &&&a_{1}&0&a_{2}\\ a_{2}&0&a_{3}&a_{5}&0&0\\ 0&a_{2}&a_{4}&a_{6}&0&0\end{pmatrix}\subset\mathbb{C}^{6}{\mathord{\otimes}}\mathbb{C}^{6},
  4. (IV)
    E=(a1a200a5a60a1000a500a1a2a3a4000a10a3a3a4a5a6000a30a500)66.E=\begin{pmatrix}a_{1}&a_{2}&0&0&-a_{5}&-a_{6}\\ 0&a_{1}&0&0&0&-a_{5}\\ 0&0&a_{1}&a_{2}&a_{3}&a_{4}\\ 0&0&0&a_{1}&0&a_{3}\\ a_{3}&a_{4}&a_{5}&a_{6}&0&0\\ 0&a_{3}&0&a_{5}&0&0\end{pmatrix}\subset\mathbb{C}^{6}{\mathord{\otimes}}\mathbb{C}^{6}.

The proof is given in §6. It utilizes classical (Atkinson-Lloyd) and algebreo-geometric techniques.

Question 2.2.

Eisenbud-Harris [12] show that once rr is large, that the basic spaces of bounded rank rr have moduli. What is the smallest rr where moduli appear?

Our new examples are part of a general construction:

Proposition 2.3.

Let EE be an 𝕒{\mathbb{a}}-dimensional space of bounded rank rr 𝐛×𝕔\mathbf{b}\times{\mathbb{c}} matrices and let FEnd(k)F\subset\operatorname{End}(\mathbb{C}^{k}) be an 𝕒{\mathbb{a}}-dimensional space of commuting matrices. Take bases of EE and FF and for each matrix of the basis of EE replace each entry with the matrix of the corresponding basis element of FF multiplied by the value of the entry to obtain a 𝐛k×𝕔k\mathbf{b}k\times{\mathbb{c}}k matrix. Call the new space E~\widetilde{E}. Then E~\widetilde{E} is a space of bounded rank at most krkr.

The proof of Proposition 2.3 is given in §5.1.

The space E~\widetilde{E} of Proposition 2.3 is interesting only if the k×kk\times k matrices are not simultaneously diagonalizable, as otherwise it is isomorphic to the direct sum of kk copies of EE.

The two new examples in the Main Theorem are when E=Λ3333E=\Lambda^{3}\mathbb{C}^{3}\subset\mathbb{C}^{3}{\mathord{\otimes}}\mathbb{C}^{3} is put in Atkinson normal form (see §5.1) and the three basis elements are respectively replaced by

(a100a1),(a200a2),(a3a5a4a6),and(a1a20a1),(a3a40a3),(a5a60a5).\begin{pmatrix}a_{1}&0\\ 0&a_{1}\end{pmatrix},\begin{pmatrix}a_{2}&0\\ 0&a_{2}\end{pmatrix},\begin{pmatrix}a_{3}&a_{5}\\ a_{4}&a_{6}\end{pmatrix},\ \ \ {\rm and}\ \ \ \begin{pmatrix}a_{1}&a_{2}\\ 0&a_{1}\end{pmatrix},\begin{pmatrix}a_{3}&a_{4}\\ 0&a_{3}\end{pmatrix},\begin{pmatrix}a_{5}&a_{6}\\ 0&a_{5}\end{pmatrix}.
Remark 2.4.

A more geometric construction that is closely related to the construction of Proposition 2.3 is presented in Proposition 3.1 below.

Remark 2.5.

Yet another method for constructing larger spaces of bounded rank from smaller ones is given in [19, §5].

We prove several technical results about invariants of spaces of bounded rank. We introduce Atkinson invariants which arise naturally in the study of Atkinson normal form for spaces of bounded rank, and show that they upper bound the first Chern classes of the sheaves introduced in [12] (Proposition 5.7). We correct a misconception in [12] about these sheaves (Proposition 5.9, Example 5.8). In the motivation described in §3.1, tensors of minimal border rank play a role. We show that such tensors cannot give rise to interesting spaces of corank one (Corollary 5.5). We exhibit new families of corank two spaces generalizing ((III)), ((IV)) in §7, as well as several variants.

Overview

In §3 we explain the correspondence between spaces of matrices and tensors, give background information on the geometry of tensors, analyze the geometry of the tensors associated to the examples in Theorem 2.1, and give several generalizations. In §4 we review the work of Atkinson-Llyod, Atkinson, and Eisenbud-Harris. In §5 we take steps to unite the linear algebra and algebraic geometry perspectives. In §6 we prove Theorem 2.1. In §7 we provide a few additional examples of spaces of bounded rank.

3. Interpretations and generalizations via the associated tensors

3.1. Background

Throughout this article, A,B,CA,B,C are complex vector spaces of dimensions 𝕒,𝐛,𝕔{\mathbb{a}},\mathbf{b},{\mathbb{c}}. We let {ai}\{a_{i}\}, {bj}\{b_{j}\}, {ck}\{c_{k}\} respectively be bases of A,B,CA,B,C. There is a 1-1 correspondence between tensors TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C, up to GL(A)×GL(B)×GL(C)GL(A)\times GL(B)\times GL(C) equivalence where the induced map TA:ABCT_{A}:A^{*}\rightarrow B{\mathord{\otimes}}C is injective, and 𝕒{\mathbb{a}}-dimensional linear subspaces of BCB{\mathord{\otimes}}C, i.e., points of the Grassmanian G(𝕒,BC)G({\mathbb{a}},B{\mathord{\otimes}}C), up to GL(B)×GL(C)GL(B)\times GL(C) equivalence, given by TT(A)T\mapsto T(A^{*}).

In the study of tensors, the tensors that are the least understood are those that are 11-degenerate: those where T(A)BCT(A^{*})\subset B{\mathord{\otimes}}C, T(B)ACT(B^{*})\subset A{\mathord{\otimes}}C and T(C)ABT(C^{*})\subset A{\mathord{\otimes}}B are of bounded rank.

A tensor TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C has (tensor) rank one if it is of the form T=abcT=a{\mathord{\otimes}}b{\mathord{\otimes}}c. The rank of TT is the smallest rr such that TT may be written as a sum of rr rank one elements, and the border rank of TT, 𝐑¯(T)\underline{\mathbf{R}}(T) is the smallest rr such that TT is a limit of tensors of rank rr. In geometric language, 𝐑¯(T)\underline{\mathbf{R}}(T) is the smallest rr such that [T]σr(Seg(A×B×C))(ABC)[T]\in\sigma_{r}(Seg(\mathbb{P}A\times\mathbb{P}B\times\mathbb{P}C))\subset\mathbb{P}(A{\mathord{\otimes}}B{\mathord{\otimes}}C), the rr-the secant variety of the Segre variety. If TA:ABCT_{A}:A^{*}\rightarrow B{\mathord{\otimes}}C is injective, then 𝐑¯(T)𝕒\underline{\mathbf{R}}(T)\geq{\mathbb{a}}. The tensor TT is called concise when all three such maps are injective. When TT is concise, if it has border rank max{𝕒,𝐛,𝕔}\operatorname{max}\{{\mathbb{a}},\mathbf{b},{\mathbb{c}}\}, one says that TT has minimal border rank.

3.2. Motivations for this project

In addition to being a classical problem of interest in its own right, two motivations for this project are as follows:

Strassen’s laser method

The problem to either unblock Strassen’s laser method for upper bounding the exponent of matrix multiplication [24], or prove it has exhausted its utility, began in [1]. The problem is to find new tensors to use in the method that can prove better upper bounds on the exponent than the big Coppersmith-Winograd tensor, or prove that no such exist. Minimal border rank 11-degenerate tensors are a class that are not yet known to have barriers to improving the method [4].

Geometric rank and cost v. value in quantum information theory

While the rank and border rank of tensors have been studied for a long time, recently attention has been paid to the notions of subrank, border subrank, and the closely related notions of slice rank [27] and geometric rank [16]. Spaces of bounded rank give rise to tensors with degenerate geometric rank. In this paper we expand upon the (at the time surprising) result of [14], where upper bounds on geometric rank imply lower bounds on border rank.

3.3. Brief discussion of tensors and their geometry

Given a tensor TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C, define its (extended) symmetry group

G^T={gGL(A)×GL(B)×GL(C)gT=T}\widehat{G}_{T}=\{g\in GL(A)\times GL(B)\times GL(C)\mid g\cdot T=T\}

and the corresponding symmetry Lie algebra 𝔤^T\widehat{\mathfrak{g}}_{T}. The actual symmetry group GTG_{T} has dimension two less as the map GL(A)×GL(B)×GL(C)GL(ABC)GL(A)\times GL(B)\times GL(C)\rightarrow GL(A{\mathord{\otimes}}B{\mathord{\otimes}}C) has a two dimensional kernel.

Several important (classes of) tensors are:

  • The unit tensor: M1mmmm=ABCM_{\langle 1\rangle}^{\oplus m}\in\mathbb{C}^{m}{\mathord{\otimes}}\mathbb{C}^{m}{\mathord{\otimes}}\mathbb{C}^{m}=A{\mathord{\otimes}}B{\mathord{\otimes}}C, where M1m=j=1majbjcjM_{\langle 1\rangle}^{\oplus m}=\sum_{j=1}^{m}a_{j}{\mathord{\otimes}}b_{j}{\mathord{\otimes}}c_{j}.

  • The WW-state, also known as a general tangent vector to the Segre variety: W=a1b1c2+a1b2c1+a2b1c1222W=a_{1}{\mathord{\otimes}}b_{1}{\mathord{\otimes}}c_{2}+a_{1}{\mathord{\otimes}}b_{2}{\mathord{\otimes}}c_{1}+a_{2}{\mathord{\otimes}}b_{1}{\mathord{\otimes}}c_{1}\in\mathbb{C}^{2}{\mathord{\otimes}}\mathbb{C}^{2}{\mathord{\otimes}}\mathbb{C}^{2}.

  • Given an algebra 𝒜{\mathcal{A}}, one may form its structure tensor T𝒜T_{{\mathcal{A}}} obtained from the bilinear map 𝒜×𝒜𝒜{\mathcal{A}}\times{\mathcal{A}}\rightarrow{\mathcal{A}} given by multiplication.

  • The matrix multiplication tensor M𝕟n2n2n2M_{\langle\mathbb{n}\rangle}\in\mathbb{C}^{n^{2}}{\mathord{\otimes}}\mathbb{C}^{n^{2}}{\mathord{\otimes}}\mathbb{C}^{n^{2}} is the special case of T𝒜T_{{\mathcal{A}}} when 𝒜{\mathcal{A}} is the algebra of n×nn\times n matrices.

  • Let GGL(A)×GL(B)×GL(C)G\subset GL(A)\times GL(B)\times GL(C) be a subgroup and let LABCL\subset A{\mathord{\otimes}}B{\mathord{\otimes}}C be a trivial GG-submodule. Any nonzero TLT\in L is a GG-invariant tensor, i.e., GG^TG\subseteq\widehat{G}_{T}. Such are particularly interesting when GG is large. However, already when GG is a regular one-dimensional torus these tensors (called tight tensors in this case) can have interesting properties. Strassen conjectures that such tensors have minimal asymptotic rank. See, e.g., [21, 8, 25].

  • A special case is when A=B=VA=B=V, C=Λ2VC=\Lambda^{2}V^{*}, G=SL(V)G=SL(V) and THom(Λ2V,VV)T\in\operatorname{Hom}(\Lambda^{2}V,V{\mathord{\otimes}}V) is the inclusion map. When dimV\operatorname{dim}V is odd, the corresponding tensor realizes two distinct spaces of bounded rank, which are cases (I), (II) of the main theorem when dimV=5\operatorname{dim}V=5. When dimV\operatorname{dim}V is even, the generalization of ((II)) is still of bounded rank.

  • A special case of the previous is when dimV=3\operatorname{dim}V=3, as then Λ2VV\Lambda^{2}V^{*}\simeq V and one obtains Tskewcw,2Λ33333T_{skewcw,2}\in\Lambda^{3}\mathbb{C}^{3}\subset\mathbb{C}^{3}{\mathord{\otimes}}\mathbb{C}^{3}{\mathord{\otimes}}\mathbb{C}^{3}.

Another motivation for this project was to find new classes of tensors of interest. In what follows we show that the tensors (III), (IV) of Theorem 2.1 have many interesting properties and generalizations.

Given TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C and TABCT^{\prime}\in A^{\prime}{\mathord{\otimes}}B^{\prime}{\mathord{\otimes}}C^{\prime}, their Kronecker product is just their tensor product considered as a three-way tensor: TT(AA)(BB)(CC)T\boxtimes T^{\prime}\in(A{\mathord{\otimes}}A^{\prime}){\mathord{\otimes}}(B{\mathord{\otimes}}B^{\prime}){\mathord{\otimes}}(C{\mathord{\otimes}}C^{\prime}).

3.4. A geometric interpretation of the tensor associated to Case (IV) and a tensor variant of Proposition 2.3

Case (IV) is a special case of the following construction:

Proposition 3.1.

Let TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C be such that T(A)T(A^{*}) has bounded rank rr and let TABCT^{\prime}\in A^{\prime}{\mathord{\otimes}}B^{\prime}{\mathord{\otimes}}C^{\prime} with 𝕒=𝐛=𝕔=m{\mathbb{a}}^{\prime}=\mathbf{b}^{\prime}={\mathbb{c}}^{\prime}=m be of minimal border rank. Then TT(AA)T\boxtimes T^{\prime}(A^{*}{\mathord{\otimes}}{A^{\prime}}^{*}) has bounded rank at most rmrm.

Proof.

The proposition clearly holds when T=M1mT^{\prime}=M_{\langle 1\rangle}^{\oplus m} has rank mm, as then one just obtains the sum of mm copies of TT. Being of bounded rank is a Zariski closed condition and all concise minimal border rank mm tensors are degenerations of M1mM_{\langle 1\rangle}^{\oplus m}, i.e., in GLm×GLm×GLmM1m¯\overline{GL_{m}\times GL_{m}\times GL_{m}\cdot M_{\langle 1\rangle}^{\oplus m}}. ∎

Case (IV) is the special case T=Tskewcw,2T=T_{skewcw,2} and T=WT^{\prime}=W.

Neither construction strictly contains the other: Case (III) does not arise from the construction of Proposition 3.1 and a minimal border rank 11-degenerate tensor does not in general correspond to a space of commuting matrices.

3.5. A geometric interpretation of Case (III) and generalizations

The matrix multiplication tensor may be defined as follows: Let U,V,WU,V,W be vector spaces and let A=UVA=U^{*}{\mathord{\otimes}}V, B=VWB=V^{*}{\mathord{\otimes}}W, C=WUC=W^{*}{\mathord{\otimes}}U. Then the (possibly rectangular) matrix multiplication tensor is the (up to scale) unique GL(U)×GL(V)×GL(W)GL(U)\times GL(V)\times GL(W)-tensor in ABCA{\mathord{\otimes}}B{\mathord{\otimes}}C, namely MU,V,W:=IdUIdVIdWM_{\langle U,V,W\rangle}:=\operatorname{Id}_{U}{\mathord{\otimes}}\operatorname{Id}_{V}{\mathord{\otimes}}\operatorname{Id}_{W}. In bases we may write M2M_{\langle 2\rangle} as

M2(A)=(x11x21x12x22x11x21x12x22).M_{\langle 2\rangle}(A^{*})=\begin{pmatrix}x^{1}_{1}&x^{1}_{2}&&\\ x^{2}_{1}&x^{2}_{2}&&\\ &&x^{1}_{1}&x^{1}_{2}\\ &&x^{2}_{1}&x^{2}_{2}\end{pmatrix}.

Consider the following construction which may be considered as an augmentation of the matrix multiplication tensor (see §7 for an even further generalization): let U,V,WU,V,W be even dimensional vector spaces equipped with symplectic forms ωU,ωV,ωW\omega_{U},\omega_{V},\omega_{W}. Let A=UVWA=U{\mathord{\otimes}}V\oplus W, B=VWUB=V{\mathord{\otimes}}W\oplus U, C=WUVC=W{\mathord{\otimes}}U\oplus V. Note that ABCA{\mathord{\otimes}}B{\mathord{\otimes}}C has a four dimensional space of Sp(U)×Sp(V)×Sp(W)Sp(U)\times Sp(V)\times Sp(W) invariant tensors with basis ωUωVωW\omega_{U}{\mathord{\otimes}}\omega_{V}{\mathord{\otimes}}\omega_{W} (which, identifying UUU\simeq U^{*} etc. using the symplectic form, is IdUIdVIdW\operatorname{Id}_{U}{\mathord{\otimes}}\operatorname{Id}_{V}{\mathord{\otimes}}\operatorname{Id}_{W}), ωUωV\omega_{U}{\mathord{\otimes}}\omega_{V}, ωVωW\omega_{V}{\mathord{\otimes}}\omega_{W}, ωUωW\omega_{U}{\mathord{\otimes}}\omega_{W}. Here Sp(U)=Sp(U,ωU)Sp(U)=Sp(U,\omega_{U}) is the symplectic group preserving ωU\omega_{U}. Let

(1) TUVW=ωUωVωW+ωUωW+ωVωWT_{UVW}=\omega_{U}{\mathord{\otimes}}\omega_{V}{\mathord{\otimes}}\omega_{W}+\omega_{U}{\mathord{\otimes}}\omega_{W}+\omega_{V}{\mathord{\otimes}}\omega_{W}

so GTUVWSp(U)×Sp(V)×Sp(W)G_{T_{UVW}}\supset Sp(U)\times Sp(V)\times Sp(W). Then when dimU=dimV=dimW=2\operatorname{dim}U=\operatorname{dim}V=\operatorname{dim}W=2 we obtain Case (III). Explicitly, Case (III) may be rewritten, setting {u1,u2}\{u_{1},u_{2}\} a basis of UUU\simeq U^{*} etc. and xji=uivjx^{i}_{j}=u_{i}{\mathord{\otimes}}v_{j},

(x11x21w2x12x22w2x11x21w1x12x22w2w1w2w1w2).\begin{pmatrix}x^{1}_{1}&x^{1}_{2}&&&-w_{2}&\\ x^{2}_{1}&x^{2}_{2}&&&&-w_{2}\\ &&x^{1}_{1}&x^{1}_{2}&w_{1}&\\ &&x^{2}_{1}&x^{2}_{2}&&w_{2}\\ w_{1}&&w_{2}&&&\\ &w_{1}&&w_{2}&&\end{pmatrix}.

This construction shows in particular that Case (III) has 2\mathbb{Z}_{2} symmetry by exchanging the BB and CC factors. Note further that the space of matrices T(B)T(B^{*}) is not of bounded rank. By the fundamental theorem of geometric rank [16] T(B)T(B^{*}) must have a highly nontransverse intersection with some σr(Seg(A×C))\sigma_{r}(Seg(\mathbb{P}A\times\mathbb{P}C)). Indeed, it intersects σ2(Seg(A×C))\sigma_{2}(Seg(\mathbb{P}A\times\mathbb{P}C)) in a 1\mathbb{P}^{1}.

Recall that the quaternion algebra is isomorphic to the algebra of 2×22\times 2 matrices, so its structure tensor is M2M_{\langle 2\rangle}. There is a six dimensional algebra 𝕊\mathbb{S} that sits between the quaternions and the octonions, called the sextonions [18]. We thank L. Manivel for observing that Case (III) is related to the sextonions.

Proposition 3.2.

The tensor TUVWT_{UVW} of (1) when dimU=dimV=dimW=2\operatorname{dim}U=\operatorname{dim}V=\operatorname{dim}W=2 is the structure tensor of the sextonions, i.e., T𝕊(A)T_{\mathbb{S}}(A^{*}) is Case (III).

Proof.

This will be more transparent if we instead examine T𝕊(C)T_{\mathbb{S}}(C^{*}) (which is isomorphic to T𝕊(B)T_{\mathbb{S}}(B^{*})). Take bases v1w1,v1w2,v2w1,v2w2,u1,u2v_{1}{\mathord{\otimes}}w_{1},v_{1}{\mathord{\otimes}}w_{2},v_{2}{\mathord{\otimes}}w_{1},v_{2}{\mathord{\otimes}}w_{2},u_{1},u_{2} of BB and u1v2,u2v2,u1v1,u2v1,w1,w2u_{1}{\mathord{\otimes}}v_{2},u_{2}{\mathord{\otimes}}v_{2},u_{1}{\mathord{\otimes}}v_{1},u_{2}{\mathord{\otimes}}v_{1},w_{1},w_{2} of AA and write yij=uiwjy_{ij}=u_{i}{\mathord{\otimes}}w_{j}. Then

(2) T(B)=(y22y21y12y11y22y21y12y11v2v1y22y12v2v1y21y11)T(B^{*})=\begin{pmatrix}y_{22}&-y_{21}&&&&\\ -y_{12}&y_{11}&&&&\\ &&-y_{22}&y_{21}&&\\ &&y_{12}&-y_{11}&&\\ &-v_{2}&&v_{1}&y_{22}&y_{12}\\ v_{2}&&-v_{1}&&y_{21}&y_{11}\end{pmatrix}

On the other hand the multiplication in 𝕊UUU\mathbb{S}\simeq U^{*}{\mathord{\otimes}}U\oplus U is given by [18, Def. 3.11] (X,μ)(X,μ)=(XX,(trace(X)IdX)μ+Xμ)(X,\mu)*(X^{\prime},\mu^{\prime})=(XX^{\prime},(\operatorname{trace}(X)\operatorname{Id}-X)\mu^{\prime}+X^{\prime}\mu). Writing the matrix representing the action of (X,μ)(X,\mu) out in bases and permuting, we obtain the result. ∎

3.6. Degeneracy loci and symmetry Lie algebras

Let T(A)BCT(A^{*})\subset B{\mathord{\otimes}}C be of bounded rank rr, where in this paragraph we allow that possibly rr is full rank. Let ΣA:={αArank(T(α))<r}\Sigma_{A}:=\{\alpha\in A^{*}\mid{\mathrm{rank}}(T(\alpha))<r\} denote the degeneracy locus of the space. Let GΣAGL(A)G_{\Sigma_{A}}\subset GL(A) denote its symmetry group and 𝔤ΣA𝔤𝔩(A)\mathfrak{g}_{\Sigma_{A}}\subset\mathfrak{g}\mathfrak{l}(A) its symmetry Lie algebra. Let πA(𝔤T)𝔤𝔩(A)\pi_{A}(\mathfrak{g}_{T})\subset\mathfrak{g}\mathfrak{l}(A) denote the projection of 𝔤T\mathfrak{g}_{T} onto the first factor. Then πA(𝔤T)𝔤ΣA\pi_{A}(\mathfrak{g}_{T})\subset\mathfrak{g}_{\Sigma_{A}}. Thus in general one obtains that

𝔤T𝔤ΣA𝔤ΣB𝔤ΣC.\mathfrak{g}_{T}\subseteq\mathfrak{g}_{\Sigma_{A}}\oplus\mathfrak{g}_{\Sigma_{B}}\oplus\mathfrak{g}_{\Sigma_{C}}.

Case (II) is a space of constant rank, and has been well-studied. Case (I) drops rank over the Grassmannian G(2,5)(Λ25)G(2,5)\subset\mathbb{P}(\Lambda^{2}\mathbb{C}^{5}). (Here and below we are interested in the projective geometry of the map so we work projectively - see §4.2 for more details.) The symmetry group of G(2,5)G(2,5) is SL5SL_{5} which is indeed the symmetry group of this tensor.

Case (IV) is 3\mathbb{Z}_{3}-invariant as it is the Kronecker product of two 3\mathbb{Z}_{3}-invariant tensors, namely Tskewcw,2WT_{skewcw,2}\boxtimes W. Here Tskewcw,2Λ33T_{skewcw,2}\in\Lambda^{3}\mathbb{C}^{3} and WS22W\in S^{2}\mathbb{C}^{2}. One has 𝔤Tskewcw,2=𝔰𝔩3\mathfrak{g}_{T_{skewcw,2}}=\mathfrak{s}\mathfrak{l}_{3} and

𝔤^W={(λ1λ120μ1ν1),(μ1μ120λ1ν1),(ν1ν120μ1λ1)λ12+μ12+ν12=0}\widehat{\mathfrak{g}}_{W}=\left\{\begin{pmatrix}\lambda_{1}&\lambda_{12}\\ 0&-\mu_{1}-\nu_{1}\end{pmatrix},\ \begin{pmatrix}\mu_{1}&\mu_{12}\\ 0&-\lambda_{1}-\nu_{1}\end{pmatrix},\ \begin{pmatrix}\nu_{1}&\nu_{12}\\ 0&-\mu_{1}-\lambda_{1}\end{pmatrix}\mid\lambda_{12}+\mu_{12}+\nu_{12}=0\right\}

Here the degeneracy locus in each factor is just a linear space, e.g. ΣA={α1=α3=α5=0}\Sigma_{A}=\{\alpha_{1}=\alpha_{3}=\alpha_{5}=0\}, so its Lie algebra in each factor is the parabolic stabilizing the three plane, which has dimension 2727. This is perhaps easier to see when we write the space as

(XYY)\begin{pmatrix}X&Y\\ Y&\end{pmatrix}

where X,YX,Y are spaces of skew-symmetric 3×33\times 3 matrices, and the degeneracy locus is when Y=0Y=0. Explicitly we have the 2121-dimensional Lie algebra

𝔤^Tskewcw,2W={(ab0a),(ab0a),(ab0a)a,b𝔰𝔩3}(Id𝔤^W).\widehat{\mathfrak{g}}_{T_{skewcw,2}\boxtimes W}=\left\{\begin{pmatrix}a&b\\ 0&a\end{pmatrix},\ \begin{pmatrix}a&b\\ 0&a\end{pmatrix},\ \begin{pmatrix}a&b\\ 0&a\end{pmatrix}\mid a,b\in\mathfrak{sl}_{3}\right\}\oplus(\mathrm{Id}\boxtimes\widehat{\mathfrak{g}}_{W}).

The large increase over 8+58+5 (where 5=dim𝔤^W5=\operatorname{dim}\widehat{\mathfrak{g}}_{W}) is striking.

Super-additivity of dimension of symmetry Lie algebras under Kronecker product fails for generic tensors: the Kronecker product of two generic tensors will have no nontrivial symmetries. For unit tensors M1kM1=M1kM_{\langle 1\rangle}^{\oplus k}\boxtimes M_{\langle 1\rangle}^{\oplus\ell}=M_{\langle 1\rangle}^{\oplus k\ell} so the (extended symmetry) jump is to 2k2k\ell compared with an expected 2k+22k+2\ell.

Question 3.3.

How to characterize situations where under Kronecker product, the dimension of the symmetry Lie algebra is super-additive?

The symmetry Lie algebra of Case (III) is 2020-dimensional, and it contains 𝔰𝔩23\mathfrak{s}\mathfrak{l}_{2}^{\oplus 3}. Explicitly, it is

𝔤^T𝕊=𝔤𝔩(U)×𝔤𝔩(V)×𝔤𝔩(W)UVW\widehat{\mathfrak{g}}_{T_{\mathbb{S}}}=\mathfrak{g}\mathfrak{l}(U)\times\mathfrak{g}\mathfrak{l}(V)\times\mathfrak{g}\mathfrak{l}(W)\oplus U{\mathord{\otimes}}V{\mathord{\otimes}}W

where on A=UVWA=U{\mathord{\otimes}}V\oplus W, the action of UVWU{\mathord{\otimes}}V{\mathord{\otimes}}W is uvw.(uv+w)=ωU(u,u)ωV(v,v)wu{\mathord{\otimes}}v{\mathord{\otimes}}w.(u^{\prime}{\mathord{\otimes}}v^{\prime}+w^{\prime})=\omega_{U}(u,u^{\prime})\omega_{V}(v,v^{\prime})w, and on B=VWUB=V{\mathord{\otimes}}W\oplus U, the action is uvw.(vw+u)=ωU(u,u)vwu{\mathord{\otimes}}v{\mathord{\otimes}}w.(v^{\prime}{\mathord{\otimes}}w^{\prime}+u^{\prime})=\omega_{U}(u,u^{\prime})v{\mathord{\otimes}}w and the action on CC is uvw.(wu+v)=ωV(v,v)uwu{\mathord{\otimes}}v{\mathord{\otimes}}w.(w^{\prime}{\mathord{\otimes}}u^{\prime}+v^{\prime})=\omega_{V}(v,v^{\prime})u{\mathord{\otimes}}w. To see UVW𝔤^T𝕊U{\mathord{\otimes}}V{\mathord{\otimes}}W\subset\widehat{\mathfrak{g}}_{T_{\mathbb{S}}}, write out T𝕊T_{\mathbb{S}} in bases

T𝕊\displaystyle T_{\mathbb{S}} =(u1v1)(v2w1)(u2w2)(u2v1)(v2w1)(u1w2)\displaystyle=(u_{1}{\mathord{\otimes}}v_{1}){\mathord{\otimes}}(v_{2}{\mathord{\otimes}}w_{1}){\mathord{\otimes}}(u_{2}{\mathord{\otimes}}w_{2})-(u_{2}{\mathord{\otimes}}v_{1}){\mathord{\otimes}}(v_{2}{\mathord{\otimes}}w_{1}){\mathord{\otimes}}(u_{1}{\mathord{\otimes}}w_{2})
(u1v2)(v1w1)(u2w2)+(u2v2)(v1w1)(u1w2)\displaystyle-(u_{1}{\mathord{\otimes}}v_{2}){\mathord{\otimes}}(v_{1}{\mathord{\otimes}}w_{1}){\mathord{\otimes}}(u_{2}{\mathord{\otimes}}w_{2})+(u_{2}{\mathord{\otimes}}v_{2}){\mathord{\otimes}}(v_{1}{\mathord{\otimes}}w_{1}){\mathord{\otimes}}(u_{1}{\mathord{\otimes}}w_{2})
(u1v1)(v2w2)(u2w1)+(u2v1)(v2w2)(u1w1)\displaystyle-(u_{1}{\mathord{\otimes}}v_{1}){\mathord{\otimes}}(v_{2}{\mathord{\otimes}}w_{2}){\mathord{\otimes}}(u_{2}{\mathord{\otimes}}w_{1})+(u_{2}{\mathord{\otimes}}v_{1}){\mathord{\otimes}}(v_{2}{\mathord{\otimes}}w_{2}){\mathord{\otimes}}(u_{1}{\mathord{\otimes}}w_{1})
+(u1v2)(v1w2)(u2w1)(u2v2)(v1w2)(u1w1)\displaystyle+(u_{1}{\mathord{\otimes}}v_{2}){\mathord{\otimes}}(v_{1}{\mathord{\otimes}}w_{2}){\mathord{\otimes}}(u_{2}{\mathord{\otimes}}w_{1})-(u_{2}{\mathord{\otimes}}v_{2}){\mathord{\otimes}}(v_{1}{\mathord{\otimes}}w_{2}){\mathord{\otimes}}(u_{1}{\mathord{\otimes}}w_{1})
+w1u1(u2w2)w2u1(u2w1)w1u2(u1w2)+w2u2(u1w1)\displaystyle+w_{1}{\mathord{\otimes}}u_{1}{\mathord{\otimes}}(u_{2}{\mathord{\otimes}}w_{2})-w_{2}{\mathord{\otimes}}u_{1}{\mathord{\otimes}}(u_{2}{\mathord{\otimes}}w_{1})-w_{1}{\mathord{\otimes}}u_{2}{\mathord{\otimes}}(u_{1}{\mathord{\otimes}}w_{2})+w_{2}{\mathord{\otimes}}u_{2}{\mathord{\otimes}}(u_{1}{\mathord{\otimes}}w_{1})
+w1(v1w2)v2w2(v1w1)v2w1(v2w2)v1+w2(v2w1)v1\displaystyle+w_{1}{\mathord{\otimes}}(v_{1}{\mathord{\otimes}}w_{2}){\mathord{\otimes}}v_{2}-w_{2}{\mathord{\otimes}}(v_{1}{\mathord{\otimes}}w_{1}){\mathord{\otimes}}v_{2}-w_{1}{\mathord{\otimes}}(v_{2}{\mathord{\otimes}}w_{2}){\mathord{\otimes}}v_{1}+w_{2}{\mathord{\otimes}}(v_{2}{\mathord{\otimes}}w_{1}){\mathord{\otimes}}v_{1}

and apply a basis vector uivjwku_{i}{\mathord{\otimes}}v_{j}{\mathord{\otimes}}w_{k} to T𝕊T_{\mathbb{S}}. One gets a sum of six terms that cancel in pairs. Here the degeneracy locus in T𝕊(A)\mathbb{P}T_{\mathbb{S}}(A^{*}) is the quadric surface {x11x22x21x12}3A\{x^{1}_{1}x^{2}_{2}-x^{1}_{2}x^{2}_{1}\}\subset\mathbb{P}^{3}\subset\mathbb{P}A^{*}. The parabolic preserving this is 12+612+6 dimensional and by the above calculation, the nilpotent part of the Lie algebra projected onto each factor is isomorphic to UVWU{\mathord{\otimes}}V{\mathord{\otimes}}W which is contained in the parabolic.

In the general case where U,V,WU,V,W are 2k2k-dimensional, one obtains a (2k)2+2k(2k)^{2}+2k dimensional space of bounded rank (2k)2(2k)^{2}. That the space is of bounded rank is transparent from the Atkinson normal form introduced below.

Question 3.4.

This construction gives rise to a series of algebras generalizing the sextonions. What properties to these algebras have? Is there interesting representation theory associated to them? (Using the sextonions, one may construct the Lie algebra 𝔢712\mathfrak{e}_{7\frac{1}{2}}.)

Remark 3.5.

In general, the Lie algebra of derivations of an algebra 𝒜{\mathcal{A}} sits inside the two factor symmetry group of its structure tensor: 𝔤^T𝒜,BC:={(Y,Z)𝔤𝔩(B)×𝔤𝔩(C)(Y,Z).T𝒜=0}\widehat{\mathfrak{g}}_{T_{{\mathcal{A}}},BC}:=\{(Y,Z)\in\mathfrak{g}\mathfrak{l}(B)\times\mathfrak{g}\mathfrak{l}(C)\mid(Y,Z).T_{{\mathcal{A}}}=0\}. The Lie algebra of derivations of the quaterions \mathbb{H} is 𝔰𝔬(3)𝔰𝔩2\mathfrak{so}(3)\cong\mathfrak{s}\mathfrak{l}_{2} and that of the sextonions is an 88-dimensional algebra containing 𝔰𝔩2\mathfrak{s}\mathfrak{l}_{2} as its semi-simple Levi factor.

3.7. Border ranks

We thank Austin Conner for providing the decompositions in the results below:

Proposition 3.6.

The tensor Tskewcw,2WT_{skewcw,2}\boxtimes W corresponding to Case (IV) has border rank nine. The following is a 3\mathbb{Z}_{3}-invariant border rank nine decomposition:

Tskewcw,2W=\displaystyle T_{skewcw,2}\boxtimes W=
limt01t63[\displaystyle\lim_{t\rightarrow 0}\frac{1}{t^{6}}\mathbb{Z}_{3}\cdot[ (a134t2a334t3a4)(23b1+tb2t3b4)(49c123tc223t2c3+23t3c4+13t5c6)\displaystyle(a_{1}-\frac{3}{4}t^{2}a_{3}-\frac{3}{4}t^{3}a_{4}){\mathord{\otimes}}(-\frac{2}{3}b_{1}+tb_{2}-t^{3}b_{4}){\mathord{\otimes}}(\frac{4}{9}c_{1}-\frac{2}{3}tc_{2}-\frac{2}{3}t^{2}c_{3}+\frac{2}{3}t^{3}c_{4}+\frac{1}{3}t^{5}c_{6})
+t2(a112ta2t4a5)(b3+tb4)(c1+12tc212t2c3+t4c5)\displaystyle+t^{2}(a_{1}-\frac{1}{2}ta_{2}-t^{4}a_{5}){\mathord{\otimes}}(-b_{3}+tb_{4}){\mathord{\otimes}}(c_{1}+\frac{1}{2}tc_{2}-\frac{1}{2}t^{2}c_{3}+t^{4}c_{5})
+(23a114t2a334t3a4+t5a6)(23b1tb2+t2b312t5b6)(23c1+tc2)].\displaystyle+(\frac{2}{3}a_{1}-\frac{1}{4}t^{2}a_{3}-\frac{3}{4}t^{3}a_{4}+t^{5}a_{6}){\mathord{\otimes}}(\frac{2}{3}b_{1}-tb_{2}+t^{2}b_{3}-\frac{1}{2}t^{5}b_{6}){\mathord{\otimes}}(\frac{2}{3}c_{1}+tc_{2})].

Here the 3\mathbb{Z}_{3} indicates cyclically permuting the factors to obtain nine terms.

Proof.

Verifying the decomposition is a routine calculation. The lower bound follows from computing the rank of the p=1p=1 Koszul flattening (see, e.g., [17, Ch. 2, §2.4]). ∎

Remark 3.7.

For any two tensors T,TT,T^{\prime} one has 𝐑¯(TT)𝐑¯(T)𝐑¯(T)\underline{\mathbf{R}}(T\boxtimes T^{\prime})\leq\underline{\mathbf{R}}(T)\underline{\mathbf{R}}(T^{\prime}). Proposition 3.6 shows 9=𝐑¯(Tskewcw,2W)<𝐑¯(Tskewcw,2)𝐑¯(W)=109=\underline{\mathbf{R}}(T_{skewcw,2}\boxtimes W)<\underline{\mathbf{R}}(T_{skewcw,2})\underline{\mathbf{R}}(W)=10. Examples where strict submultiplicativity of border rank under Kronecker product are of interest and also potentially useful for proving upper bounds on the exponent of matrix multiplication. In particular Tskewcw,2T_{skewcw,2} could potentially be used to prove the exponent is two, and strict submultiplicativity of its Kronecker square has already been observed [9]. We are currently examining which other tensors exhibit strict submultiplicativity under Kronecker product with WW to potentially advance the laser method.

Proposition 3.8.

The tensor T𝕊T_{\mathbb{S}} associated to Case (III) has border rank 1010.

Proof.

The upper bound comes from the following decomposition, which is easily verified:

T𝕊=\displaystyle T_{\mathbb{S}}=
limt01t5[\displaystyle\lim_{t\rightarrow 0}\frac{1}{t^{5}}[ (a1t2a3+t4a6)(tb2+t3b3+t4b4tb5+b6)(tc4c6)\displaystyle(a_{1}-t^{2}a_{3}+t^{4}a_{6}){\mathord{\otimes}}(tb_{2}+t^{3}b_{3}+t^{4}b_{4}-tb_{5}+b_{6}){\mathord{\otimes}}(tc_{4}-c_{6})
+\displaystyle+ (a1+ta2)(t4b4+b6)(t3c1+12t2c3+c6)\displaystyle(a_{1}+ta_{2}){\mathord{\otimes}}(t^{4}b_{4}+b_{6}){\mathord{\otimes}}(-t^{3}c_{1}+\frac{1}{2}t^{2}c_{3}+c_{6})
+\displaystyle+ (a1+t2a3)(t2b1+tb2+t3b3+b6)(t4c2+tc4tc5c6)\displaystyle(-a_{1}+t^{2}a_{3}){\mathord{\otimes}}(t^{2}b_{1}+tb_{2}+t^{3}b_{3}+b_{6}){\mathord{\otimes}}(-t^{4}c_{2}+tc_{4}-tc_{5}-c_{6})
+\displaystyle+ (a1+t4a6)(tb5b6)(t3c1t4c212t2c3c6)]\displaystyle(-a_{1}+t^{4}a_{6}){\mathord{\otimes}}(tb_{5}-b_{6}){\mathord{\otimes}}(t^{3}c_{1}-t^{4}c_{2}-\frac{1}{2}t^{2}c_{3}-c_{6})]
+1t4[\displaystyle+\frac{1}{t^{4}}[ (a1+ta2+t3a5)b5(12tc5+c6)\displaystyle(-a_{1}+ta_{2}+t^{3}a_{5}){\mathord{\otimes}}b_{5}{\mathord{\otimes}}(-\frac{1}{2}tc_{5}+c_{6})
+\displaystyle+ a2(tb5+b6)(t3c1+12tc5c6)\displaystyle a_{2}{\mathord{\otimes}}(tb_{5}+b_{6}){\mathord{\otimes}}(t^{3}c_{1}+\frac{1}{2}tc_{5}-c_{6})
+\displaystyle+ (a1t2a3+t3a5)(tb1+b5)(t3c1t2c3+tc4c6)\displaystyle(a_{1}-t^{2}a_{3}+t^{3}a_{5}){\mathord{\otimes}}(tb_{1}+b_{5}){\mathord{\otimes}}(t^{3}c_{1}-t^{2}c_{3}+tc_{4}-c_{6})
\displaystyle- (a1t2a3+t3a4)(tb2+b6)c5]\displaystyle(a_{1}-t^{2}a_{3}+t^{3}a_{4}){\mathord{\otimes}}(tb_{2}+b_{6}){\mathord{\otimes}}c_{5}]
+1t3[\displaystyle+\frac{1}{t^{3}}[ (a1t3a5)(tb1t2b312b5)(2t3c2tc3+c5)\displaystyle(a_{1}-t^{3}a_{5}){\mathord{\otimes}}(-tb_{1}-t^{2}b_{3}-\frac{1}{2}b_{5}){\mathord{\otimes}}(2t^{3}c_{2}-tc_{3}+c_{5})
+\displaystyle+ (a2+2t2a4)(t3b3+12b6)(2t3c2+tc3+c5)].\displaystyle(-a_{2}+2t^{2}a_{4}){\mathord{\otimes}}(-t^{3}b_{3}+\frac{1}{2}b_{6}){\mathord{\otimes}}(-2t^{3}c_{2}+tc_{3}+c_{5})].

To obtain the lower bound we use the border substitution method [20, 22]. Note already from Strassen’s commutation conditions applied to T𝕊(B)T_{\mathbb{S}}(B^{*}), after changing bases so one basis element is the identity and using it to identify T𝕊(B)T_{\mathbb{S}}(B^{*}) as a space of endomorphisms, the commutator has full rank which implies 𝐑¯(T𝕊)9\underline{\mathbf{R}}(T_{\mathbb{S}})\geq 9. The border substitution method says that if for every Borel fixed hyperplane the commutator still has full rank, one obtains an improvement of one in the estimate. Here there are two Borel fixed hyperplanes, obtained respectively by setting v2=0v_{2}=0 and y22=0y_{22}=0. In both cases the commutator is still of full rank. ∎

4. Review of previous work

4.1. Work of Atkinson-Llyod and Atkinson [3, 2]

Given TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C with E=T(A)BCE=T(A^{*})\subset B{\mathord{\otimes}}C of bounded rank, to make connections with existing literature we write, for αA\alpha\in A^{*}, ϕ(α)=ϕT(α):BC\phi(\alpha)=\phi_{T}(\alpha):B^{*}\rightarrow C for the induced map.

Consider compression spaces of bounded rank rr of Example 1.1: they may be described invariantly as: there exist BBB^{\prime}\subset B^{*} of codimension k1k_{1}, CCC^{\prime}\subset C of dimension k2k_{2}, with k1+k2=rk_{1}+k_{2}=r, such that T(A)(B)CT(A^{*})(B^{\prime})\subset C^{\prime}. This notion is symmetric as in this case T(A)(C)BT(A^{*})({C^{\prime}}^{\perp})\subset{B^{\prime}}^{\perp}.

Since compression spaces are “understood”, one would like to eliminate them from the study as well as spaces that are sums of a compression space with another space. Hence the following definition: A bounded rank rr space is imprimitive if there exists HBH\subset B^{*} such that ϕ|H\phi|_{H} is of bounded rank r1r-1 or if there exists HCH\subset C^{*} such that ϕ𝕥|H\phi^{\mathbb{t}}|_{H} is of bounded rank r1r-1. In this case the bounded rank condition is inherited from the smaller space. Otherwise it is primitive.

Thus to classify spaces of bounded rank, it suffices to classify the primitive ones.

Atkinson and Llyod [3] showed that for primitive spaces of bounded rank rr in BCB{\mathord{\otimes}}C, with 𝐛𝕔\mathbf{b}\leq{\mathbb{c}}, either 𝐛=r+1\mathbf{b}=r+1 and 𝕔(r+12){\mathbb{c}}\leq\binom{r+1}{2} or there exist r1,r22r_{1},r_{2}\geq 2 with r1+r2=rr_{1}+r_{2}=r, 𝐛r1+1+(r2+12)\mathbf{b}\leq r_{1}+1+\binom{r_{2}+1}{2} and 𝕔r2+1+(r1+12){\mathbb{c}}\leq r_{2}+1+\binom{r_{1}+1}{2}.

Thus to classify primitive spaces of bounded rank 44, it suffices to classify them in 66\mathbb{C}^{6}{\mathord{\otimes}}\mathbb{C}^{6}, and 5𝕔\mathbb{C}^{5}{\mathord{\otimes}}\mathbb{C}^{\mathbb{c}}, with 5𝕔105\leq{\mathbb{c}}\leq 10.

4.2. Work of Eisenbud and Harris [12]

Given TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C, with T(A)T(A^{*}) of bounded rank rr, Eisenbud and Harris observed that one gets a map between vector bundles

(3) B¯\displaystyle\underline{B}^{*} C¯𝒪A(1)\displaystyle{\mathord{\;\longrightarrow\;}}\underline{C}{\mathord{\otimes}}{\mathcal{O}}_{\mathbb{P}A^{*}}(1)
\displaystyle\searrow \displaystyle\ \ \ \ \swarrow
A\displaystyle\mathbb{P}A^{*}

The notation is that C¯:=C𝒪A\underline{C}:=C{\mathord{\otimes}}{\mathcal{O}}_{\mathbb{P}A^{*}} is the trivial vector bundle with fiber CC. Let =A{\mathcal{E}}={\mathcal{E}}_{A} denote the image sheaf and =A{\mathcal{F}}={\mathcal{F}}_{A} the image sheaf of the map with the roles of B,CB,C reversed. Since T(A)T(A^{*}) has bounded rank rr, both sheaves locally free off of a codimension at least two subset because ,{\mathcal{E}},{\mathcal{F}} are torsion free. We slightly abuse notation, writing ϕ\phi for the horizontal map (3).

They show that if T(A)T(A^{*}) is primitive and c1()=1c_{1}({\mathcal{E}})=1, then the space is obtained as a projection of the classical Example 1.3. I.e., BAB\subseteq A and CΛ2AC\subseteq\Lambda^{2}A^{*}.

They assert (1){\mathcal{E}}^{**}\cong{\mathcal{F}}^{*}(1) and by symmetry (1){\mathcal{F}}^{**}\cong{\mathcal{E}}^{*}(1), but this is false in general, see Example 5.11. Their assertion would imply c1(A)+c1(A)=rc_{1}({\mathcal{E}}_{A})+c_{1}({\mathcal{F}}_{A})=r, which they do not assert but instead just assert c1(A)+c1(A)rc_{1}({\mathcal{E}}_{A})+c_{1}({\mathcal{F}}_{A})\leq r, which we verify and identify the failure of equality to hold (Proposition 5.9).

Remark 4.1.

When one has a space of constant rank the assertion is correct as will be clear by Proposition 5.9 and moreover in this case (1){\mathcal{E}}\cong{\mathcal{F}}^{*}(1) as when {\mathcal{E}} is a vector bundle, {\mathcal{E}}\cong{\mathcal{E}}^{**}.

In particular, to classify the first open case of r=4r=4, it suffices to understand when c1()=c1()=2c_{1}({\mathcal{E}})=c_{1}({\mathcal{F}})=2.

They remark that the only basic r=4r=4 case they know of with c1()=c1()=2c_{1}({\mathcal{E}})=c_{1}({\mathcal{F}})=2 is BC=5B\cong C=\mathbb{C}^{5} and A=Λ2BBBA=\Lambda^{2}B\subset B{\mathord{\otimes}}B.

We always have BH0()B^{*}\subseteq H^{0}({\mathcal{E}}) and CH0()C^{*}\subseteq H^{0}({\mathcal{F}}). Explicitly, the first inclusion is β([α]T(α,β)α\beta\mapsto([\alpha]\mapsto T(\alpha,\beta){\mathord{\otimes}}\alpha^{*}.

Eisenbud and Harris point out that imprimitivity is equivalent to {\mathcal{E}} or {\mathcal{F}} having a trivial summand. To see this, take a complement LL to HBH\subset B^{*} and consider ϕ|L:L\phi|_{L}:L\rightarrow{\mathcal{E}}. It is injective as ϕ\phi has bounded rank r>r1r>r-1, giving the desired splitting.

Note that since H0()¯\underline{H^{0}({\mathcal{E}})} surjects onto {\mathcal{E}}, we can equally well have a trivial quotient of {\mathcal{E}}, i.e., a surjection 𝒪{\mathcal{E}}\rightarrow{\mathcal{O}} as this induces a surjection H0()𝒪H^{0}({\mathcal{E}})\rightarrow{\mathcal{O}} which is just a linear map between vector spaces giving the splitting of {\mathcal{E}}. But this in turn is equivalent to having an inclusion 𝒪{\mathcal{O}}\rightarrow{\mathcal{E}}^{*}, i.e., that h0()>0h^{0}({\mathcal{E}}^{*})>0.

In summary:

Proposition 4.2.

EE is primitive if and only if h0()=h0()=0h^{0}({\mathcal{E}}^{*})=h^{0}({\mathcal{F}}^{*})=0.

To further eliminate redundancies such as EE being a subspace of a primitive space or a projection of a larger primitive space, Eisenbud and Harris defined a vector space of bounded rank T(A)T(A^{*}) to be basic if

  1. (1)

    it is strongly indecomposable: {\mathcal{E}} and {\mathcal{F}} are indecomposable. This is is a strengthening of primitivity.

  2. (2)

    it is unexpandable: φ\varphi and φ(1)\varphi^{*}(1) are the linear parts of kernels of maps of graded free [A]\mathbb{C}[A^{*}] modules. This says the space is not a projection of a space of bounded rank in some larger space of matrices.

  3. (3)

    it is unliftable: it is not a proper subspace of a family of the same rank in Hom(B,C)\mathrm{Hom}(B^{*},C).

Draisma [10] gives a sufficient condition for a space to be unliftable (he calls unliftability rank criticality): for EHom(B,C)E\subset Hom(B^{*},C) a linear space of morphisms of generic rank rr, define the space of rank neutral directions

RND(E):\displaystyle RND(E): ={XHom(B,C),X(Ker(Y))Im(Y)YE,rank(Y)=r}\displaystyle=\{X\in Hom(B^{*},C),\;X(Ker(Y))\subset Im(Y)\;\forall Y\in E,rank(Y)=r\}
=YE,rank(Y)=rT^Yσr(Seg(B×C)).\displaystyle=\bigcap_{Y\in E,rank(Y)=r}\widehat{T}_{Y}\sigma_{r}(Seg(\mathbb{P}B\times\mathbb{P}C)).

Here σr(Seg(B×C))\sigma_{r}(Seg(\mathbb{P}B\times\mathbb{P}C)) denotes the variety of rank at most rr elements and T^Y\widehat{T}_{Y} its affine tangent space at YY.

Proposition 4.3.

[10] RND(E)RND(E) always contains EE and in case of equality, EE is rank critical.

5. Towards uniting the linear algebra and algebraic geometry perspectives

5.1. Atkinson normal form

Lemma 5.1.

[2, Lemma 3] Let E𝐛𝕔E\subset\mathbb{C}^{\mathbf{b}}{\mathord{\otimes}}\mathbb{C}^{\mathbb{c}} be of bounded rank rr and suppose we have chosen bases so that some matrix X1EX_{1}\in E is of the form

X1=(Idr000).X_{1}=\begin{pmatrix}\operatorname{Id}_{r}&0\\ 0&0\end{pmatrix}.

Then for any XEX\in E we have, with the same blocking,

(4) X=(𝕩WU0)X=\begin{pmatrix}{\mathbb{x}}&W\\ U&0\end{pmatrix}

and U𝕩kW=0U{\mathbb{x}}^{k}W=0 for all k0k\geq 0.

Note that it is sufficient to check U𝕩kW=0U{\mathbb{x}}^{k}W=0 for 0kr10\leq k\leq r-1 because higher powers of 𝕩{\mathbb{x}} may be expressed as linear combinations of 𝕩0,,𝕩r1{\mathbb{x}}^{0},\dotsc,{\mathbb{x}}^{r-1}.

Proof.

That XX has the bottom right block equal to zero follows by considering appropriate minors of X+εX1X+\varepsilon X_{1}. Consider X1+tXX_{1}+tX and the size r+1r+1 minor

0=det(Idr+t𝕩tWjtUk0)0=\operatorname{det}\begin{pmatrix}\operatorname{Id}_{r}+t{\mathbb{x}}&tW_{j}\\ tU^{k}&0\end{pmatrix}

where WjW_{j} is the jj-th column of WW and UkU^{k} is the kk-th row of UU. Using the formula for a block determinant and assuming tt is sufficiently small,

(5) det(Idr+t𝕩tWjtUk0)=t2Uk(Idr+t𝕩)Wj1=t2Uk(s=0(1)sts𝕩s)Wj.\operatorname{det}\begin{pmatrix}\operatorname{Id}_{r}+t{\mathbb{x}}&tW_{j}\\ tU^{k}&0\end{pmatrix}=t^{2}U^{k}(\operatorname{Id}_{r}+t{\mathbb{x}}){}^{-1}W_{j}=t^{2}U^{k}(\sum_{s=0}^{\infty}(-1)^{s}t^{s}{\mathbb{x}}^{s})W_{j}.

The coefficient of each power of tt must vanish. Since this holds for all j,kj,k, we conclude. ∎

We add the observation that the normal form is sufficient as well:

Proposition 5.2.

Let E𝐛𝕔E\subset\mathbb{C}^{\mathbf{b}}{\mathord{\otimes}}\mathbb{C}^{\mathbb{c}} admit a presentation in Atkinson normal form (4). Then EE is of bounded rank rr.

Proof.

The normal form assures all size r+1r+1 minors containing the upper-left r×rr\times r block are zero. Say there were a size r+1r+1 minor having rank ss in its UU block and tt in its WW block. We may normalize

U=(Ids000),W=(00Idt0),U=\begin{pmatrix}\operatorname{Id}_{s}&0\\ 0&0\end{pmatrix},W=\begin{pmatrix}0&0\\ \operatorname{Id}_{t}&0\end{pmatrix},

where the blockings are respectively (s,rs)×(s,𝐛rs)(s,r-s)\times(s,\mathbf{b}-r-s) and (t,𝕔rt)×(rt,t)(t,{\mathbb{c}}-r-t)\times(r-t,t). The UW=0UW=0 equation implies the first ss rows of WW and the last tt columns of UU are zero, so s+trs+t\leq r. Write r+1=s+t+ur+1=s+t+u. There must be a size uu minor in the 𝕩{\mathbb{x}} block contributing, and that minor must be in the upper right (rs)×(rt)(r-s)\times(r-t) block of 𝕩{\mathbb{x}}. But the U𝕩W=0U{\mathbb{x}}W=0 equation implies that the 𝕩{\mathbb{x}} block must be zero in the first ss rows of this upper right block and the last tt columns of this block. So the contribution must lie in a block of size (rst)×(rst)(r-s-t)\times(r-s-t). But u=rst+1u=r-s-t+1, a contradiction. ∎

Proof of Proposition 2.3.

Put EE in Atkinson normal form. Then E~\widetilde{E} will also be in Atkinson normal form when the linear form times Idr\operatorname{Id}_{r} is replaced by a rank kk linear form. If not, just add in such a multiple of Idr\operatorname{Id}_{r} to put the space in the normal form, as this will not effect the U~𝕩~jW~=0\widetilde{U}\widetilde{\mathbb{x}}^{j}\widetilde{W}=0 equations. ∎

Atkinson essentially observes that the equations U𝕩jW=0U{\mathbb{x}}^{j}W=0 may be encoded as

(6) (UU𝕩U𝕩2U𝕩r1)(W𝕩W𝕩2W𝕩r1W)=(0),\begin{pmatrix}U\\ U{\mathbb{x}}\\ U{\mathbb{x}}^{2}\\ \vdots\\ U{\mathbb{x}}^{r-1}\end{pmatrix}\begin{pmatrix}W&{\mathbb{x}}W&{\mathbb{x}}^{2}W&\cdots&{\mathbb{x}}^{r-1}W\end{pmatrix}=(0),

where (0)(0) is (𝕔r)×(𝐛r)({\mathbb{c}}-r)\times(\mathbf{b}-r). Call the first matrix on the left AtLAt_{L} and call the second AtRAt_{R}. Recall that if X,YX,Y are respectively p×rp\times r and r×qr\times q matrices, with p,qrp,q\geq r and rank(XY)s{\mathrm{rank}}(XY)\leq s then dimLker(X)+dimRkerYrs\operatorname{dim}\operatorname{Lker}(X)+\operatorname{dim}\operatorname{Rker}Y\geq r-s, so in our case we have (rrank(X))+(rrank(Y))r(r-{\mathrm{rank}}(X))+(r-\operatorname{rank}(Y))\geq r, i.e. rank(X)+rank(Y)r{\mathrm{rank}}(X)+{\mathrm{rank}}(Y)\leq r, with XX the first matrix and YY the second.

Definition 5.3.

Notations as above. Assume bases have been chosen generically to have the normal form. Write atLat_{L} for the rank of AtLAt_{L} and atRat_{R} for the rank of AtRAt_{R} and call these the Atkinson numbers associated to EE.

In particular we have

(7) atL+atRr.at_{L}+at_{R}\leq r.

Say EBCE\subset B{\mathord{\otimes}}C has bounded rank rr and we fix X1=(Idr000)X_{1}=\begin{pmatrix}\operatorname{Id}_{r}&0\\ 0&0\end{pmatrix}, i.e., we take a general point of EE. X1:BCX_{1}:B^{*}\rightarrow C defines subspaces ImHom(B,C)(X1)C\operatorname{Im}_{\operatorname{Hom}(B^{*},C)}(X_{1})\subset C and kerHom(B,C)(X1)=ImHom(C,B)(X1)B\operatorname{ker}_{\operatorname{Hom}(B^{*},C)}(X_{1})^{\perp}=\operatorname{Im}_{\operatorname{Hom}(C^{*},B)}(X_{1})\subset B. Call these subspaces C1,B1C_{1},B_{1}. Let XEX\in E be another general element. Consider X:BCmodC1X:B^{*}\rightarrow C\operatorname{mod}C_{1}, and X:CBmodB1X:C^{*}\rightarrow B\operatorname{mod}B_{1}. The normal form implies that X(B1)C1X(B_{1}^{\perp})\subseteq C_{1} and X(C1)B1X(C_{1}^{\perp})\subseteq B_{1} The ranks of these maps give two additional invariants. In the coordinate expressions, these are respectively the maps given by UU and WW. We now allow XX to vary, and we ask what is the dimension of the subspaces of (B1)C1(B_{1}^{\perp})^{*}{\mathord{\otimes}}C_{1} and (C1)B1(C_{1}^{\perp})^{*}{\mathord{\otimes}}B_{1} that are filled out? These give two additional invariants, call them dU,dWd_{U},d_{W}. Let dUWd_{UW} be the dimension of the combined subspaces, in bases, the number of variables appearing in U,WU,W, combined. We know by conciseness (to avoid trivialities) that (B1)C1(B_{1}^{\perp})^{*}{\mathord{\otimes}}C_{1} and (C1)B1(C_{1}^{\perp})^{*}{\mathord{\otimes}}B_{1} respectively surject onto C1,B1C_{1},B_{1} which lower bounds their dimensions.

Proposition 5.4.

If r=𝕔1r={\mathbb{c}}-1 and dU=1d_{U}=1 or r=𝐛1r=\mathbf{b}-1 and dW=1d_{W}=1 then EE is imprimitive. More generally, for any 𝐛,𝕔,r\mathbf{b},{\mathbb{c}},r, if UU can be normalized to be zero except in one column or WW can be normalized to be zero except in one row, then EE is imprimitive.

Proof.

Say r=𝐛1r=\mathbf{b}-1 and dW=1d_{W}=1. Then in Atkinson normal form we may normalize W=(w1,0,,0)𝕥W=(w_{1},0,\dotsc,0)^{\mathbb{t}}, so

E=(w10U0)E=\begin{pmatrix}*&w_{1}\\ *&0\\ U&0\end{pmatrix}

where the blocking is (𝐛1,1)×(1,𝐛2,𝕔𝐛+1)(\mathbf{b}-1,1)\times(1,\mathbf{b}-2,{\mathbb{c}}-\mathbf{b}+1). If we strike out the first row the corank is unchanged. In the general case, the same argument holds when

E=(w1wk00U00).E=\begin{pmatrix}*&w_{1}&\cdots&w_{k}\\ *&0&\cdots&0\\ U&0&\cdots&0\end{pmatrix}.

Corollary 5.5.

Assume 𝕒𝐛=𝕔{\mathbb{a}}\leq\mathbf{b}={\mathbb{c}}. A concise tensor TABCT\in A{\mathord{\otimes}}B{\mathord{\otimes}}C such that T(A)T(A^{*}) is corank one and primitive cannot have minimal border rank 𝐛\mathbf{b}.

Proof.

By [15, Prop. 3.3], a tensor of minimal border rank as in the hypothesis must have dU=dW=1d_{U}=d_{W}=1. ∎

We often slightly abuse notation and let U,𝕩,WU,{\mathbb{x}},W denote the matrices of linear forms induced by EE rather than individual matrices.

Let EE be primitive and let Ann(W)\text{Ann}\,(W) denote the submodule of length rr row vectors with entries polynomials in the variables appearing in WW that annihilate WW.

Proposition 5.6.

Let EBCE\subset B{\mathord{\otimes}}C be concise of bounded rank rr and primitive. If Ann(W)\text{Ann}\,(W) is generated by row vectors of linear forms and atL=dim(Ann(W))=𝐛r+1at_{L}=\operatorname{dim}(\text{Ann}\,(W))=\mathbf{b}-r+1, where dim(Ann(W))\operatorname{dim}(\text{Ann}\,(W)) is the dimension of the vector space spanned by minimal generators of Ann(W)\text{Ann}\,(W), then EE is expandable.

Proof.

Put EE in Atkinson normal (𝕩WU0)\begin{pmatrix}{\mathbb{x}}&W\\ U&0\end{pmatrix}. Write uαu_{\alpha} for the rows of UU. Let

(AtL)prim=uα𝕩j1α𝐛r,j0Ann(W).(At_{L})_{prim}=\langle u_{\alpha}{\mathbb{x}}^{j}\mid 1\leq\alpha\leq\mathbf{b}-r,j\geq 0\rangle\subseteq\text{Ann}\,(W).

Let uminimal generators of Ann(W)\uα1α𝐛ru\in\langle\textrm{minimal generators of }\text{Ann}\,(W)\rangle\backslash\langle u_{\alpha}\mid 1\leq\alpha\leq\mathbf{b}-r\rangle. Then E~=(𝕩WU0u0)\widetilde{E}=\begin{pmatrix}{\mathbb{x}}&W\\ U&0\\ u&0\end{pmatrix} is also of bounded rank rr by Proposition 5.2 as it satisfies the conditions of Atkinson normal form. ∎

5.2. Chern classes and Atkinson numbers

Proposition 5.7.

Notations as in §4.2, then c1()=atRc1(τR)c_{1}({\mathcal{E}})=at_{R}-c_{1}(\tau_{R}) and c1()=atLc1(τL)c_{1}({\mathcal{F}})=at_{L}-c_{1}(\tau_{L}), where τR,τL\tau_{R},\tau_{L} are torsion sheafs that are described explicitly in the proof. In particular, c1()atRc_{1}({\mathcal{E}})\leq at_{R} and c1()atLc_{1}({\mathcal{F}})\leq at_{L}.

Proof of Proposition 5.7.

Put the matrix of φ:B¯C¯(1)\varphi:\underline{B}^{*}\to\underline{C}(1) in Atkinson normal form:

(φ𝕩φWφU0).\begin{pmatrix}\varphi_{\mathbb{x}}&\varphi_{W}\\ \varphi_{U}&0\end{pmatrix}.

Over the point [α]V[\alpha]\in\mathbb{P}V^{*} where ϕ[α]=(Idr000)\phi_{[\alpha]}=\begin{pmatrix}\operatorname{Id}_{r}&0\\ 0&0\end{pmatrix}, let B2BB^{*}_{2}\subset B^{*} denote ker(ϕ[α])\operatorname{ker}(\phi_{[\alpha]}) and let B1B^{*}_{1} be a complement in BB^{*}. In other words, B¯1\underline{B}^{*}_{1} is the trivial subbundle of B¯\underline{B}^{*} generated by the first rr basis vectors. Similarly write C=C1C2C=C_{1}\oplus C_{2}. Note that we may identify B1B^{*}_{1} and C1C_{1}.

We analyze c1()c_{1}({\mathcal{E}}). Consider the following maps:

ϕi:B¯2(1)B¯2(2)B¯2(i)\displaystyle\phi_{i}\quad\colon\quad\underline{B}^{*}_{2}(-1)\oplus\underline{B}^{*}_{2}(-2)\oplus\cdots\oplus\underline{B}^{*}_{2}(-i) C¯1\displaystyle\rightarrow\underline{C}_{1}
t1ti\displaystyle t_{1}\oplus\cdots\oplus t_{i} φW(t1)+φ𝕩φW(t2)++φ𝕩i1φW(ti).\displaystyle\mapsto\varphi_{W}(t_{1})+\varphi_{\mathbb{x}}\varphi_{W}(t_{2})+\cdots+\varphi_{\mathbb{x}}^{i-1}\varphi_{W}(t_{i}).

Note that Im(ϕi)=Im(ϕi+1)Im(\phi_{i})=Im(\phi_{i+1}) for all iri\geq r. Denote the image of ϕr\phi_{r} by 𝒦1{\mathcal{K}}_{1} and cokernel of ϕr\phi_{r} by 𝒬1\mathcal{Q}_{1}. Then rank𝒦1=atR\operatorname{rank}{\mathcal{K}}_{1}=at_{R}. Explicitly, we may write 𝒦1{\mathcal{K}}_{1} as a subsheaf of B¯1\underline{B}^{*}_{1} as follows: Let 𝒱A\mathcal{V}\subset\mathbb{P}A^{*} be an affine open subset containing [α][\alpha]. Using the identification B1C1B_{1}^{*}\cong C_{1},

𝒦1𝒱={sH0(B1𝒱)tiH0(B¯2(i)𝒱) such that s=φW(t1)+φ𝕩φW(t2)++φ𝕩r1φW(tr)}.{\mathcal{K}}_{1}\mid_{\mathcal{V}}=\{s\in H^{0}(B^{*}_{1}\mid_{\mathcal{V}})\ \mid\exists t_{i}\in H^{0}(\underline{B}^{*}_{2}(-i)\mid_{\mathcal{V}})\text{\ such\ that\ }s=\varphi_{W}(t_{1})+\varphi_{\mathbb{x}}\varphi_{W}(t_{2})+\ldots+\varphi_{\mathbb{x}}^{r-1}\varphi_{W}(t_{r})\}.

Since for all k0k\geq 0 and all tB2t\in B_{2}^{*}, φUφ𝕩kφW(t)=0\varphi_{U}\varphi_{\mathbb{x}}^{k}\varphi_{W}(t)=0, we obtain φ(𝒦1)𝒦1(1)C¯1(1)\varphi({\mathcal{K}}_{1})\subseteq{\mathcal{K}}_{1}(1)\subset\underline{C}_{1}(1) and rankφ𝒦1=rank𝒦1=atR\operatorname{rank}\varphi\mid_{{\mathcal{K}}_{1}}=\operatorname{rank}{\mathcal{K}}_{1}=at_{R}. We have the following commutative diagram that defines φ𝒬\varphi^{\mathcal{Q}}:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒦1\textstyle{{\mathcal{K}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ𝒦1\scriptstyle{\varphi\mid_{{\mathcal{K}}_{1}}}B¯1B¯2\textstyle{\underline{B}^{*}_{1}\oplus\underline{B}^{*}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}𝒬1B¯2\textstyle{\mathcal{Q}_{1}\oplus\underline{B}^{*}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ𝒬\scriptstyle{\varphi^{\mathcal{Q}}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒦1(1)\textstyle{{\mathcal{K}}_{1}(1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C¯1(1)C¯2(1)\textstyle{\underline{C}_{1}(1)\oplus\underline{C}_{2}(1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒬1(1)C¯2(1)\textstyle{\mathcal{Q}_{1}(1)\oplus\underline{C}_{2}(1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Since φ𝒦1\varphi\mid_{{\mathcal{K}}_{1}} is of full rank, ker(φ𝒦1)\operatorname{ker}(\varphi\mid_{{\mathcal{K}}_{1}}) is torsion, but since it is a subsheaf of the torsion free sheaf B¯1\underline{B}_{1}^{*}, it is zero, and coker(φ𝒦1)\mathrm{coker}(\varphi\mid_{{\mathcal{K}}_{1}}) is a torsion sheaf. The first Chern classes in the short exact sequence 0𝒦1𝒦1(1)coker(φ𝒦1)00\rightarrow{\mathcal{K}}_{1}\rightarrow{\mathcal{K}}_{1}(1)\rightarrow\text{coker}(\varphi\mid_{{\mathcal{K}}_{1}})\rightarrow 0 show

c1(coker(φ𝒦1))=rank𝒦1=atR.c_{1}(\text{coker}(\varphi\mid_{{\mathcal{K}}_{1}}))=\operatorname{rank}{\mathcal{K}}_{1}=at_{R}.

Since coker(φ𝒦1)(\varphi\mid_{{\mathcal{K}}_{1}}) is torsion, any subsheaf of it has first Chern class no larger than atRat_{R}.

Claim: c1(kerφ𝒬)0c_{1}(\ker\varphi^{\mathcal{Q}})\geq 0.

To see this, since φ(B¯2)=φW(B¯2)𝒦1(1)\varphi(\underline{B}^{*}_{2})=\varphi_{W}(\underline{B}^{*}_{2})\subseteq{\mathcal{K}}_{1}(1), we have B¯2kerφ𝒬\underline{B}^{*}_{2}\subseteq\ker\varphi^{\mathcal{Q}}. The kernel of the map φ𝒬𝒬1:𝒬1𝒬1(1)C¯2(1)\varphi^{\mathcal{Q}}\mid_{\mathcal{Q}_{1}}\colon\mathcal{Q}_{1}\rightarrow\mathcal{Q}_{1}(1)\oplus\underline{C}_{2}(1) is torsion because φ𝒬𝒬1\varphi^{\mathcal{Q}}\mid_{\mathcal{Q}_{1}} is of full rank (since φB1\varphi\mid_{B_{1}^{*}} is of full rank). Hence c1(kerφ𝒬𝒬1)0c_{1}(\ker\varphi^{\mathcal{Q}}\mid_{\mathcal{Q}_{1}})\geq 0. Together we have c1(kerφ𝒬)0c_{1}(\ker\varphi^{\mathcal{Q}})\geq 0.

Hence we have c1(Im(φ𝒬))=c1(𝒬1B¯2)c1(kerφ𝒬)=c1(𝒬1)c1(kerφ𝒬)c_{1}(\operatorname{Im}(\varphi^{\mathcal{Q}}))=c_{1}(\mathcal{Q}_{1}\oplus\underline{B}^{*}_{2})-c_{1}(\ker\varphi^{\mathcal{Q}})=c_{1}(\mathcal{Q}_{1})-c_{1}(\ker\varphi^{\mathcal{Q}}).

We also have the following commutative diagram:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ𝒦1\textstyle{\ker\varphi\mid_{{\mathcal{K}}_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ\textstyle{\ker\varphi\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ𝒬\textstyle{\ker\varphi^{\mathcal{Q}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cokerφ𝒦1\textstyle{\mathrm{coker}\varphi\mid_{{\mathcal{K}}_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cokerφ\textstyle{\mathrm{coker}\varphi\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cokerφ𝒬\textstyle{\mathrm{coker}\varphi^{\mathcal{Q}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒦1\textstyle{{\mathcal{K}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B¯1B¯2\textstyle{\underline{B}^{*}_{1}\oplus\underline{B}^{*}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒬1B¯2\textstyle{\mathcal{Q}_{1}\oplus\underline{B}^{*}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

whose rows are exact.

Define 𝒜:=ker(cokerφ𝒦1cokerφ){\mathcal{A}}\colon=\ker(\mathrm{coker}\varphi\mid_{{\mathcal{K}}_{1}}\to\mathrm{coker}\varphi) and let {\mathcal{B}} denote the corresponding quotient. We have

0𝒜cokerφ𝒦10.0\longrightarrow{\mathcal{A}}\longrightarrow\mathrm{coker}\varphi\mid_{{\mathcal{K}}_{1}}\longrightarrow{\mathcal{B}}\longrightarrow 0.

Recall that kerφ𝒦1=0\ker\varphi\mid_{{\mathcal{K}}_{1}}=0. Consider the exact sequence

0kerφkerφ𝒬𝒜0.0\longrightarrow\ker\varphi\longrightarrow\ker\varphi^{\mathcal{Q}}\longrightarrow{\mathcal{A}}\longrightarrow 0.

Thus c1()=c1(kerφ)=c1(kerφ𝒬)+c1(𝒜)=atRc1(ker(φ𝒬))c1()c_{1}({\mathcal{E}})=-c_{1}(\ker\varphi)=-c_{1}(\ker\varphi^{\mathcal{Q}})+c_{1}({\mathcal{A}})=at_{R}-c_{1}(\operatorname{ker}(\varphi^{\mathcal{Q}}))-c_{1}({\mathcal{B}}).

The case of c1()c_{1}({\mathcal{F}}) is similar. ∎

Example 5.8.

Let T(A)BCT(A^{*})\subset B{\mathord{\otimes}}C correspond to a maximal dimension (k1,k2)(k_{1},k_{2}) compression space of rank r=k1+k2r=k_{1}+k_{2} so that 𝕒=𝐛k1+𝕔k2k1k2{\mathbb{a}}=\mathbf{b}k_{1}+{\mathbb{c}}k_{2}-k_{1}k_{2}. A standard way to write such a space is (0)\begin{pmatrix}*&*\\ *&0\end{pmatrix} where the blocking is (k1,𝐛k1)×(k2,𝕔k2)(k_{1},\mathbf{b}-k_{1})\times(k_{2},{\mathbb{c}}-k_{2}). Permute basis vectors to put it in Atkinson normal form

(8) (𝐳1𝐳2𝐳30𝐳400𝐳50)\begin{pmatrix}\mathbf{z}_{1}&\mathbf{z}_{2}&\mathbf{z}_{3}\\ 0&\mathbf{z}_{4}&0\\ 0&\mathbf{z}_{5}&0\end{pmatrix}

where the blocking is (k1,k2,𝐛r)×(k1,k2,𝕔r)(k_{1},k_{2},\mathbf{b}-r)\times(k_{1},k_{2},{\mathbb{c}}-r). Then

(UU𝕩U𝕩2U𝕩r1)=((0,𝐳5)(0,𝐳5)(𝐳1𝐳20𝐳3)(0,𝐳5)(𝐳1𝐳20𝐳3)2)=((0,𝐳5)(0,𝐳5𝐳3))\begin{pmatrix}U\\ U{\mathbb{x}}\\ U{\mathbb{x}}^{2}\\ \vdots\\ U{\mathbb{x}}^{r-1}\end{pmatrix}=\begin{pmatrix}(0,\mathbf{z}_{5})\\ (0,\mathbf{z}_{5})\begin{pmatrix}\mathbf{z}_{1}&\mathbf{z}_{2}\\ 0&\mathbf{z}_{3}\end{pmatrix}\\ (0,\mathbf{z}_{5})\begin{pmatrix}\mathbf{z}_{1}&\mathbf{z}_{2}\\ 0&\mathbf{z}_{3}\end{pmatrix}^{2}\\ \vdots\end{pmatrix}=\begin{pmatrix}(0,\mathbf{z}_{5})\\ (0,\mathbf{z}_{5}\mathbf{z}_{3})\\ \vdots\end{pmatrix}

Since the entries of 𝐳5\mathbf{z}_{5} are independent, we obtain atL=k2at_{L}=k_{2}. Since det(𝐳1)\operatorname{det}(\mathbf{z}_{1}) will factor out of the matrix of size rr minors from the first rr columns of (8), but the remaining minors are independent because the entries of 𝐳4,𝐳5\mathbf{z}_{4},\mathbf{z}_{5} are independent, we also get c1()=k2c_{1}({\mathcal{E}})=k_{2}. Similarly atR=c1()=k1at_{R}=c_{1}({\mathcal{F}})=k_{1}. If we specialize to a subspace of A\mathbb{P}A^{*}, then the Atkinson numbers and Chern classes may drop.

5.3. Chern class defects

Consider

(9) 0C¯(1)𝒬00\rightarrow{\mathcal{E}}\rightarrow\underline{C}(1)\rightarrow\mathcal{Q}_{{\mathcal{E}}}\rightarrow 0

and the analogous 𝒬\mathcal{Q}_{{\mathcal{F}}}. We sometimes write 𝒬=𝒬\mathcal{Q}=\mathcal{Q}_{\mathcal{E}} in what follows.

Proposition 5.9.

Notations as in §4.2, then c1()+c1()=rc1(xt1(𝒬,𝒪A))c_{1}({\mathcal{E}})+c_{1}({\mathcal{F}})=r-c_{1}(\mathcal{E}xt^{1}(\mathcal{Q}_{{\mathcal{E}}},\mathcal{O}_{\mathbb{P}A^{*}})). In particular c1(xt1(𝒬,𝒪A))=c1(xt1(𝒬,𝒪A))c_{1}(\mathcal{E}xt^{1}(\mathcal{Q}_{{\mathcal{E}}},\mathcal{O}_{\mathbb{P}A^{*}}))=c_{1}(\mathcal{E}xt^{1}(\mathcal{Q}_{{\mathcal{F}}},\mathcal{O}_{\mathbb{P}A^{*}})).

Note that c1(xt1(𝒬,𝒪A))0c_{1}(\mathcal{E}xt^{1}(\mathcal{Q},\mathcal{O}_{\mathbb{P}A^{*}}))\geq 0 because it is torsion.

We thank M. Popa for suggesting Proposition 5.9 and an outline of the proof.

Proof of Proposition 5.9.

Dualize (9) to get

(10) 0𝒬C¯(1)xt1(𝒬,𝒪A).0\rightarrow\mathcal{Q}^{*}\rightarrow\underline{C}^{*}(-1)\rightarrow{\mathcal{E}}^{*}\rightarrow\mathcal{E}xt^{1}(\mathcal{Q},\mathcal{O}_{\mathbb{P}A^{*}}).

Now ϕ(1):C¯B¯(1)\phi^{*}(1):\underline{C}^{*}\rightarrow\underline{B}(1) is the map giving rise to {\mathcal{F}}, so ϕ:C¯(1)B¯\phi^{*}:\underline{C}^{*}(-1)\rightarrow\underline{B} has image (1){\mathcal{F}}(-1).

Claim: we also have a map B¯{\mathcal{E}}^{*}\rightarrow\underline{B}. To see this, take a resolution of {\mathcal{E}} by vector bundles

𝒱2𝒱1B¯0\mathcal{V}_{2}\rightarrow\mathcal{V}_{1}\rightarrow\underline{B}^{*}\rightarrow{\mathcal{E}}\rightarrow 0

and dualize to get

B¯𝒱1𝒱2{\mathcal{E}}^{*}\rightarrow\underline{B}\rightarrow\mathcal{V}_{1}^{*}\rightarrow\mathcal{V}_{2}^{*}

which also gives

0𝒬C¯(1)ϕB¯𝒱1𝒱20\rightarrow\mathcal{Q}^{*}\rightarrow\underline{C}^{*}(-1)\smash{\mathop{\longrightarrow}\limits^{\phi^{*}}}\underline{B}\rightarrow\mathcal{V}_{1}^{*}\rightarrow\mathcal{V}_{2}^{*}

which is not in general exact at B¯\underline{B}. By definition, the homology at that step is xt1(𝒬,𝒪A){\mathcal{E}}xt^{1}(\mathcal{Q},{\mathcal{O}}_{\mathbb{P}A^{*}}).

By (10), ϕ\phi^{*} factors through the map B¯{\mathcal{E}}^{*}\rightarrow\underline{B}. We get an exact sequence

(11) 0(1)xt1(𝒬,𝒪A)00\rightarrow{\mathcal{F}}(-1)\rightarrow{\mathcal{E}}^{*}\rightarrow\mathcal{E}xt^{1}(\mathcal{Q},\mathcal{O}_{\mathbb{P}A^{*}})\rightarrow 0

Thus c1()=c1()r+c1(xt1(𝒬,𝒪A))c_{1}({\mathcal{E}}^{*})=c_{1}({\mathcal{F}})-r+c_{1}(\mathcal{E}xt^{1}(\mathcal{Q},\mathcal{O}_{\mathbb{P}A^{*}})), i.e.,

c1()+c1()=rc1(xt1(𝒬,𝒪A)).c_{1}({\mathcal{E}})+c_{1}({\mathcal{F}})=r-c_{1}(\mathcal{E}xt^{1}(\mathcal{Q},\mathcal{O}_{\mathbb{P}A^{*}})).

Remark 5.10.

Dualizing (11), we see the error in the assertion about {\mathcal{E}}^{**} and (1){\mathcal{F}}^{*}(1) is measured by

0(1)xt1(xt1(𝒬,𝒪A),𝒪A))xt1(,𝒪A).0\rightarrow{\mathcal{E}}^{**}\rightarrow{\mathcal{F}}^{*}(1)\rightarrow\mathcal{E}xt^{1}(\mathcal{E}xt^{1}(\mathcal{Q},\mathcal{O}_{\mathbb{P}A^{*}}),\mathcal{O}_{\mathbb{P}A^{*}}))\rightarrow\mathcal{E}xt^{1}({\mathcal{E}}^{*},\mathcal{O}_{\mathbb{P}A^{*}})\rightarrow\ldots.
Example 5.11.

Consider the following spaces from [15], which, as tensors form a complete list of concise, 11-degenerate tensors in (5)3(\mathbb{C}^{5})^{{\mathord{\otimes}}3} of minimal border rank:

Represented as spaces of matrices, the tensors may be presented as:

T𝒪58\displaystyle T_{{\mathcal{O}}_{58}} =(x1x2x3x5x5x1x4x2x1x5x1x5),T𝒪57=(x1x2x3x5x1x4x2x1x1x5),\displaystyle=\begin{pmatrix}x_{1}&&x_{2}&x_{3}&x_{5}\\ x_{5}&x_{1}&x_{4}&-x_{2}&\\ &&x_{1}&&\\ &&-x_{5}&x_{1}&\\ &&&x_{5}&\end{pmatrix},\ \ T_{{\mathcal{O}}_{57}}=\begin{pmatrix}x_{1}&&x_{2}&x_{3}&x_{5}\\ &x_{1}&x_{4}&-x_{2}&\\ &&x_{1}&&\\ &&&x_{1}&\\ &&&x_{5}&\end{pmatrix},
T𝒪56\displaystyle T_{{\mathcal{O}}_{56}} =(x1x2x3x5x1+x5x4x1x1x5),T𝒪55=(x1x2x3x5x1x5x4x1x1x5),T𝒪54=(x1x2x3x5x1x4x1x1x5).\displaystyle=\begin{pmatrix}x_{1}&&x_{2}&x_{3}&x_{5}\\ &x_{1}+x_{5}&&x_{4}&\\ &&x_{1}&&\\ &&&x_{1}&\\ &&&x_{5}&\end{pmatrix},\ \ T_{{\mathcal{O}}_{55}}=\begin{pmatrix}x_{1}&&x_{2}&x_{3}&x_{5}\\ &x_{1}&x_{5}&x_{4}&\\ &&x_{1}&&\\ &&&x_{1}&\\ &&&x_{5}&\end{pmatrix},\ \ T_{{\mathcal{O}}_{54}}=\begin{pmatrix}x_{1}&&x_{2}&x_{3}&x_{5}\\ &x_{1}&&x_{4}&\\ &&x_{1}&&\\ &&&x_{1}&\\ &&&x_{5}&\end{pmatrix}.

These are all bounded rank four. The Chern classes (c1(),c1())(c_{1}({\mathcal{E}}),c_{1}({\mathcal{F}})) equal the Atkinson numbers and are respectively (2,2),(1,1),(1,1),(1,1),(1,1)(2,2),(1,1),(1,1),(1,1),(1,1). This is most easily computed via the Chern classes of the kernel bundles.

These are all compression spaces, as permuting the third and fifth columns makes the lower 3×33\times 3 block zero. They are degenerations of the general such compression space over 15\mathbb{P}^{15} where the kernel sheaves are isomorphic and move in a Seg(1×2)5Seg(\mathbb{P}^{1}\times\mathbb{P}^{2})\subset\mathbb{P}^{5}, which means 𝒦=𝒪15(2){\mathcal{K}}={\mathcal{O}}_{\mathbb{P}^{15}}(-2). In particular the Chern classes can decrease under specialization.

Note in the (1,1)(1,1) cases we have c1()=1c_{1}({\mathcal{E}})=1, c1()=c1((1))=41=3c_{1}({\mathcal{E}}^{**})=c_{1}({\mathcal{F}}^{*}(1))=4-1=3, so the Eisenbud-Harris assertions about {\mathcal{E}} and (1){\mathcal{F}}^{*}(1) fail in these cases.

6. Proof of the Main Theorem

Outline of the proof: In §6.1, we make general remarks on the c1()=2c_{1}({\mathcal{E}})=2 corank one case. The (5,5)(5,5) case is treated in §6.2. The (5,m)(5,m) cases, with m>7m>7 are ruled out by [2, Thm. B]. We refine the proof of [2, Thm. B] to rule out the (5,7)(5,7) case in §6.3. We reformulate the Kronecker normal form for pencils of matrices and use it to analyze linear annihilators of spaces of linear forms of 2×r2\times r matrices in §6.4. In §6.5 we use our analysis in §6.4 to reduce to two new potentially basic spaces. The (6,5)(6,5) case is treated in §6.6. In §6.7 we go to a slightly more general setting and observe the Chern classes in the two (6,6)(6,6) examples are indeed all 22. In §6.8 we prove the two (6,6)(6,6) examples are indeed basic.

6.1. Zero sets and syzygies of spaces of quadrics

Continuing the notation of §4.2, write

0𝒦B¯C(1),0\rightarrow{\mathcal{K}}_{\mathcal{E}}\rightarrow\underline{B}^{*}\rightarrow C(1),

and

0𝒦C¯B(1),0\rightarrow{\mathcal{K}}_{\mathcal{F}}\rightarrow\underline{C}^{*}\rightarrow B(1),

which we dualize and twist to get

B¯C(1)𝒦(1).\underline{B}^{*}\rightarrow C(1)\rightarrow{\mathcal{K}}_{\mathcal{F}}^{*}(1).

Since the maps BC(1)B^{*}\rightarrow C(1) agree we obtain

(12) 0𝒦B¯C(1)𝒦(1),0\rightarrow{\mathcal{K}}_{\mathcal{E}}\rightarrow\underline{B}^{*}\rightarrow C(1)\rightarrow{\mathcal{K}}_{\mathcal{F}}^{*}(1),

which is a complex but not in general exact. Write the maps as d3,d2,d1d_{3},d_{2},d_{1} respectively.

Recall the dictionary between sheaves on projective space and graded modules. Write R=[A]R=\mathbb{C}[A] for the ring of polynomials on AA^{*} and adopt the usual convention that R(k)R(k) is RR with the grading shifted by kk. In terms of modules, assuming c1()=c1()=2c_{1}({\mathcal{E}})=c_{1}({\mathcal{F}})=2, the corank one (m,m)(m,m) case becomes:

𝔽:\textstyle{\mathbb{F}_{\bullet}\quad\colon}R(2)\textstyle{R(-2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ\scriptstyle{\ker\varphi}Rm\textstyle{R^{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}Rm(1)\textstyle{R^{m}(1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ𝕥\scriptstyle{\ker\varphi^{\mathbb{t}}}R(3).\textstyle{R(3).}

Here the entries of the vector kerφ\operatorname{ker}\varphi are quadrics, which, by hypothesis have no common factor, and similarly for the entries of kerφ𝕥\operatorname{ker}\varphi^{\mathbb{t}}.

In this section we reduce the corank one case with c1()=2c_{1}({\mathcal{E}})=2. We already saw that the kernel map is given by a row vector of quadrics which have no common factor. We then specialize to the basic rank four case and show the zero set of the quadrics always has codimension at least three.

For BS2AB\subset S^{2}A^{*}, let K21(B)K_{21}(B) be the space of linear syzygies, i.e., the kernel of the multiplication map BAS3AB{\mathord{\otimes}}A^{*}\rightarrow S^{3}A^{*}. We will be interested in the corank one space of matrices given by the tensor inducing ABK21(B)=:BCA^{*}\subset B{\mathord{\otimes}}K_{21}(B)^{*}=:B{\mathord{\otimes}}C.

Let BS2AB\subset S^{2}A^{*}, let 𝐛=dimB3\mathbf{b}=\operatorname{dim}B\geq 3 (the case of 𝐛=2\mathbf{b}=2 is handled by Kronecker normal form).

Let Q1,Q2BQ_{1},Q_{2}\in B be general and consider their zero set. Then either

  1. (1)

    The zero set is a degree four codimension two irreducible complete intersection. In this case intersecting with a third general quadric in BB will provide a codimension three set.

  2. (2)

    The zero set is of pure codimension two, consisting of the union of a cone over the twisted cubic v3(1)v_{3}(\mathbb{P}^{1}) and a linear space.

  3. (3)

    The zero set is codimension two and the codimension two component consists of two irreducible quadrics, each in a hyperplane. In this case the space spanned by Q1,Q2Q_{1},Q_{2} is Q,m\langle Q,\ell m\rangle, where QQ is irreducible and ,mA\ell,m\in A^{*}, and the two irreducible quadrics are Zeros(Q,){\rm Zeros}(Q,\ell), Zeros(Q,m){\rm Zeros}(Q,m). Then, by the genericity hypothesis, B=Q,FB=\langle Q,\ell F\rangle where dimF=𝐛1\operatorname{dim}F=\mathbf{b}-1 and dimK21(B)=(𝐛12)\operatorname{dim}K_{21}(B)=\binom{\mathbf{b}-1}{2}.

  4. (4)

    The zero set is codimension two and the codimension two components consist of linear spaces. Then B=1F1,2F2B=\langle\ell_{1}F_{1},\ell_{2}F_{2}\rangle, for some linear subspaces F1,F2AF_{1},F_{2}\subset A^{*}.

  5. (5)

    The zero set has codimension one: then B=FB=\ell F where A\ell\in A^{*} and FAF\subset A^{*} and then dimK21(B)=(𝐛2)\operatorname{dim}K_{21}(B)=\binom{\mathbf{b}}{2}.

The only case requiring explanation is (2), which follows from the classification of varieties of minimal degree, see, e.g., [11], which in particular says that a degree three irreducible variety of codimension two is a cone over the twisted cubic.

We now consider what possible basic spaces of bounded rank 𝐛1\mathbf{b}-1 in BC=BK21(B)B{\mathord{\otimes}}C=B{\mathord{\otimes}}K_{21}(B) could arise with kernel of dimension one given by the quadrics in BB. We assume 𝐛>4\mathbf{b}>4. We claim only case (1) can potentially occur: Case (2) cannot occur because in this case dimK21(B)=2\operatorname{dim}K_{21}(B)=2. The remaining cases are ruled out because in each case there will be vectors of linear forms in the kernel.

6.2. Conclusion of the (5,5)(5,5) case

Lemma 6.1.

The only basic space of 5×55\times 5 matrices of linear forms with c1()=c1()=2c_{1}({\mathcal{E}})=c_{1}({\mathcal{F}})=2 is the space of skew-symmetric matrices.

Proof.

Let II denote the ideal generated by the quadrics in kerφ\operatorname{ker}\varphi. By the discussion in §6.1, depthI=codimI3\mathrm{depth}\;I=\mathrm{codim}\;I\geq 3.

Set Ik(φ)I_{k}(\varphi) to be the ideal generated by the size kk minors of ϕ\phi. Using that Λ4ϕ=(kerϕ)(kerϕ𝕥)\langle\Lambda^{4}\phi\rangle=(\operatorname{ker}\phi)(\operatorname{ker}\phi^{\mathbb{t}}), we have gcd(I4(φ))=1\mathrm{gcd}(I_{4}(\varphi))=1 as c1()=c1()=2c_{1}({\mathcal{E}})=c_{1}({\mathcal{F}})=2. This implies depthI4(φ)2\mathrm{depth}\;I_{4}(\varphi)\geq 2 and depthI1(kerφ)3\mathrm{depth}\;I_{1}(\ker\varphi)\geq 3. This is enough for cokerφ\mathrm{coker}\;\varphi to be a first syzygy module, so that we can complete the resolution 𝔽\mathbb{F}_{\bullet} on the left:

𝔽:0\textstyle{\mathbb{F}_{\bullet}\quad\colon\quad 0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}R(2)\textstyle{R(-2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ\scriptstyle{\ker\varphi}R5\textstyle{R^{5}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}R5(1)\textstyle{R^{5}(1)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kerφ𝕥\scriptstyle{\ker\varphi^{\mathbb{t}}}R(3)\textstyle{R(3)}

By assumption, since φ\varphi is primitive, the entries in kerφ𝕥\ker\varphi^{\mathbb{t}} are linearly independent. We are forced to have depthI1(kerφ𝕥)=codimI1(kerφ𝕥)3\mathrm{depth}\;I_{1}(\ker\varphi^{\mathbb{t}})=\mathrm{codim}\;I_{1}(\ker\varphi^{\mathbb{t}})\geq 3. By the Buchsbaum-Eisenbud criterion for exactness [5], we conclude that both 𝔽\mathbb{F}_{\bullet} and its dual 𝔽\mathbb{F}_{\bullet}^{*} are exact. Hence 𝔽\mathbb{F}_{\bullet} is self dual (up to shift) and it defines a Gorenstein ideal of depth 33. By the characterization of Gorenstein ideals of depth 3 [6, Thm. 2.1(2)], φ\varphi is skew-symmetrizable. ∎

6.3. (r+1)×m(r+1)\times m spaces of bounded rank rr

A basis of the linear annihilator of a vector of linear forms (a1,,ak)(a_{1},\dotsc,a_{k}) is

(a2a10),(a20a10),,(ak00a1),(0a3a20),(0a40a20),,(0ak00a2),,(00akak1).\begin{pmatrix}-a_{2}\\ a_{1}\\ 0\\ \vdots\\ \vdots\end{pmatrix},\begin{pmatrix}-a_{2}\\ 0\\ a_{1}\\ 0\\ \vdots\end{pmatrix},\ldots,\begin{pmatrix}-a_{k}\\ 0\\ \vdots\\ 0\\ a_{1}\end{pmatrix},\begin{pmatrix}0\\ -a_{3}\\ a_{2}\\ 0\\ \vdots\\ \vdots\end{pmatrix},\begin{pmatrix}0\\ -a_{4}\\ 0\\ a_{2}\\ 0\\ \vdots\end{pmatrix},\ldots,\begin{pmatrix}0\\ -a_{k}\\ 0\\ \vdots\\ 0\\ a_{2}\end{pmatrix},\ldots,\begin{pmatrix}0\\ \vdots\\ 0\\ -a_{k}\\ a_{k-1}\end{pmatrix}.

In particular, it has dimension (k2)\binom{k}{2}.

From this, one may recover the Atkinson result:

Theorem 6.2.

[2, Thm. B] Let Er+1mE\subset\mathbb{C}^{r+1}{\mathord{\otimes}}\mathbb{C}^{m} be a primitive space of bounded rank rr and let m>r+(r12)m>r+\binom{r-1}{2}. Then EE is a specialization of r+1Hom(r+1,Λ2r+1)\mathbb{C}^{r+1}\subset\operatorname{Hom}(\mathbb{C}^{r+1},\Lambda^{2}\mathbb{C}^{r+1}).

The idea of the proof is that assuming UU has no entries equal to zero, there are enough elements of the linear annihilator of UU present so the U𝕩W=0U{\mathbb{x}}W=0 equation forces 𝕩{\mathbb{x}} to be a scalar times the identity. Writing out the matrix of linear forms of VHom(V,Λ2V)V\rightarrow\operatorname{Hom}(V,\Lambda^{2}V) one sees this implies the space is a specialization of VHom(V,Λ2V)V\rightarrow\operatorname{Hom}(V,\Lambda^{2}V).

We now consider rank 44 spaces in 57\mathbb{C}^{5}{\mathord{\otimes}}\mathbb{C}^{7}:

First assume all entries of U=(u1,u2,u3,u4)U=(u_{1},u_{2},u_{3},u_{4}) are linearly independent. In order to avoid 𝕩{\mathbb{x}} being forced to be a linear form times the identity by the U𝕩W=0U{\mathbb{x}}W=0 equation, WW is uniquely determined up to isomorphism, namely

W=(u2u3u4u1u1u1).W=\begin{pmatrix}-u_{2}&-u_{3}&-u_{4}\\ u_{1}&&\\ &u_{1}&\\ &&u_{1}\end{pmatrix}.

Here u12u_{1}^{2} divides all size three minors, and one would obtain c1()=1c_{1}({\mathcal{F}})=1, so this case is eliminated.

When UU is three dimensional, normalize so U=(0,u2,u3,u4)U=(0,u_{2},u_{3},u_{4}). Then if the first row of WW is zero, the full annihilator of (u2,u3,u4)(u_{2},u_{3},u_{4}) must appear in WW. Then using the U𝕩W=0U{\mathbb{x}}W=0 equation we see that the first row of 𝕩{\mathbb{x}} is zero except in the (1,1)(1,1) slot. Otherwise the first row of WW contains a nonzero entry and U𝕩W=0U{\mathbb{x}}W=0 gives the same conclusion for the first column of 𝕩{\mathbb{x}}. Striking the first row (resp. column) gives a space of bounded rank r1r-1 so the space is not primitive.

When UU is two-dimensional, write U=(0,0,u3,u4)U=(0,0,u_{3},u_{4}), so we must have, after a change of basis

W=(w11w21w31w12w22w32u4ϵu3ϵ).W=\begin{pmatrix}w^{1}_{1}&w^{1}_{2}&w^{1}_{3}\\ w^{2}_{1}&w^{2}_{2}&w^{2}_{3}\\ &&-u_{4}\epsilon\\ &&u_{3}\epsilon\end{pmatrix}.

with the upper left 2×22\times 2 block of full rank to avoid a column of zeros after a change of basis and ϵ{0,1}\epsilon\in\{0,1\}. Then U𝕩W=0U{\mathbb{x}}W=0 implies the lower left 2×22\times 2 block of 𝕩{\mathbb{x}} is zero. Combining this with the first two columns of WW and below, we obtain a 4×34\times 3 block of zeros and the space is compression.

When UU is one-dimensional, the space is imprimitive by Proposition 5.4.

We conclude:

Proposition 6.3.

Let E57E\subset\mathbb{C}^{5}{\mathord{\otimes}}\mathbb{C}^{7} be a primitive space of bounded rank 44. Then EE is a specialization of 5Hom(5,Λ25)\mathbb{C}^{5}\subset\operatorname{Hom}(\mathbb{C}^{5},\Lambda^{2}\mathbb{C}^{5}).

6.4. 2×k2\times k spaces of linear forms and their linear annihilators

Recall the Kronecker normal form for pencils of matrices, i.e., tensors in ABFA{\mathord{\otimes}}B{\mathord{\otimes}}F with dimF=2\operatorname{dim}F=2: all are block matrices with blocks

Lk(F):=(st000st0st),Lk𝕥(F)=(s000ts00ts),Jork,λ(F)=sIdk+tJL_{k}(F^{*}):=\begin{pmatrix}s&t&0&\cdots&0\\ 0&s&t&0&\cdots&\\ &&\ddots&&\\ &&&s&t\end{pmatrix},\ L_{k}^{\mathbb{t}}(F^{*})=\begin{pmatrix}s&0&0&\cdots&0\\ t&s&0&0&\cdots&\\ &&\ddots&&\\ &&\ddots&&\\ &&&t&s\end{pmatrix},\ Jor_{k,\lambda}(F^{*})=s\operatorname{Id}_{k}+tJ

where the matrices are respectively (k+1)×k(k+1)\times k, k×(k+1)k\times(k+1), k×kk\times k and in the last case JJ is a single Jordan block with eigenvalue λ\lambda. Of particular note is L1(F)=(s,t)L_{1}(F^{*})=(s,t) and L1𝕥(F)=(st)L_{1}^{\mathbb{t}}(F^{*})=\begin{pmatrix}s\\ t\end{pmatrix}. We adopt the convention that the rows are indexed by AA and the columns by BB. Rewritten as a linear subspace of BFB{\mathord{\otimes}}F these become:

Lk(A)=(a1a2ak00a1ak),Lk𝕥(A)=(a1a2aka2a3ak+1),L_{k}(A^{*})=\begin{pmatrix}a_{1}&a_{2}&\cdots&a_{k}&0\\ 0&a_{1}&&\cdots&a_{k}\end{pmatrix},\ L_{k}^{\mathbb{t}}(A^{*})=\begin{pmatrix}a_{1}&a_{2}&\cdots&a_{k}\\ a_{2}&a_{3}&\cdots&a_{k+1}\end{pmatrix},
Jork,λ(A)=(a1a2akλa1λa2+a1λak+ak1).Jor_{k,\lambda}(A^{*})=\begin{pmatrix}a_{1}&a_{2}&\cdots&a_{k}\\ \lambda a_{1}&\lambda a_{2}+a_{1}&\cdots&\lambda a_{k}+a_{k-1}\end{pmatrix}.

Note the special cases

L1(A)=(a00a),L1𝕥(A)=(a1a2),Jor1,0(A)=(a0),Jor1,λ(A)=(aλa),Jor2,0=(a1a20a1).L_{1}(A^{*})=\begin{pmatrix}a&0\\ 0&a\end{pmatrix},\ L_{1}^{\mathbb{t}}(A^{*})=\begin{pmatrix}a_{1}\\ a_{2}\end{pmatrix},\ Jor_{1,0}(A^{*})=\begin{pmatrix}a\\ 0\end{pmatrix},\ Jor_{1,\lambda}(A^{*})=\begin{pmatrix}a\\ \lambda a\end{pmatrix},\ Jor_{2,0}=\begin{pmatrix}a_{1}&a_{2}\\ 0&a_{1}\end{pmatrix}.

Let 0uvf0\leq u\leq v\leq f, Write A=A1AfA=A_{1}\oplus\cdots\oplus A_{f}. The general form is,

(Lk1(A1),,Lku(Au),L1(Au+1),,Lv(Av),Jori1,λi1(Av+1),,Jorif,λif(Af)).(L_{k_{1}}(A_{1}^{*}),\dotsc,L_{k_{u}}(A_{u}^{*}),L_{\ell_{1}}(A_{u+1}^{*}),\dotsc,L_{\ell_{v}}(A_{v}^{*}),Jor_{i_{1},\lambda_{i_{1}}}(A_{v+1}^{*}),\dotsc,Jor_{i_{f},\lambda_{i_{f}}}(A_{f}^{*})).

Now we study linear annihilators. First note that if a space consists of a single block, it has no linear annihilator. We next consider pairs of blocks. Label the linear forms in the first block with aja_{j}’s and those in the second with bjb_{j}’s.

An LkL_{k} and an LqL_{q}, or an LkL_{k} and a Jq,0J_{q,0} with q>1q>1, or a Jk,0J_{k,0}, k>1k>1 and a Jq,0J_{q,0}, q>1q>1: one gets a two dimensional space. Respectively,

(b100a100),(00bq00aq),and(b100a100),(b2b10a1a20),and(b1000a1000),(b2b100a2a100).\begin{pmatrix}-b_{1}\\ 0\\ \vdots\\ 0\\ a_{1}\\ 0\\ \vdots\\ 0\end{pmatrix},\begin{pmatrix}0\\ \vdots\\ 0\\ -b_{q}\\ 0\\ \vdots\\ 0\\ a_{q}\end{pmatrix},\ \ {\rm and}\ \ \begin{pmatrix}-b_{1}\\ 0\\ \vdots\\ 0\\ a_{1}\\ 0\\ \vdots\\ 0\end{pmatrix},\begin{pmatrix}-b_{2}\\ -b_{1}\\ 0\\ \vdots\\ a_{1}\\ a_{2}\\ 0\\ \vdots\end{pmatrix},\ \ {\rm and}\ \ \begin{pmatrix}-b_{1}\\ 0\\ 0\\ \vdots\\ 0\\ a_{1}\\ 0\\ 0\\ \vdots\\ 0\end{pmatrix},\begin{pmatrix}-b_{2}\\ -b_{1}\\ 0\\ \vdots\\ 0\\ a_{2}\\ a_{1}\\ 0\\ \vdots\\ 0\end{pmatrix}.

An LkL_{k} with a J1,0J_{1,0} has a one-dimensional linear annihilator. A Jk,0J_{k,0} (any kk) with a J1,0J_{1,0} also has a one-dimensional linear annihilator.

An L1L_{1} and an Lq𝕥L_{q}^{\mathbb{t}} or a Jorq,λJor_{q,\lambda}: one gets a qq dimensional space. Respectively:

(b1b200a100),,(bqbq+10000a1),and(b1λb100a100),(b2λb2b3000a10),,(bqλbqbq+10000a1).\begin{pmatrix}-b_{1}\\ -b_{2}\\ 0\\ \vdots\\ 0\\ a_{1}\\ 0\\ \vdots\\ 0\end{pmatrix},\dotsc,\begin{pmatrix}-b_{q}\\ -b_{q+1}\\ 0\\ \vdots\\ 0\\ 0\\ 0\\ \vdots\\ a_{1}\end{pmatrix},\ \ {\rm and}\ \ \begin{pmatrix}-b_{1}\\ -\lambda b_{1}\\ 0\\ \vdots\\ 0\\ a_{1}\\ 0\\ \vdots\\ 0\end{pmatrix},\begin{pmatrix}-b_{2}\\ -\lambda b_{2}-b_{3}\\ 0\\ \vdots\\ 0\\ 0\\ a_{1}\\ 0\\ \vdots\end{pmatrix},\dotsc,\begin{pmatrix}-b_{q}\\ -\lambda b_{q}-b_{q+1}\\ 0\\ \vdots\\ 0\\ 0\\ 0\\ \vdots\\ a_{1}\end{pmatrix}.

All other pairs have no linear annihilator.

6.5. (6,6)(6,6)-spaces

We now specialize to spaces of 2×42\times 4 matrices with linear forms as entries with at least a two-dimensional linear annihilator. The discussion in §6.4 implies the linear annihilator is at most two-dimensional.

The same argument as in the (5,7)(5,7) case shows that UU cannot have any columns equal to zero.

Going through partitions of four, we already know we need at least two parts.

The case of (3,1)(3,1) is ruled out by our general analysis above.

We consider the (2,2)(2,2) and (2,1,1)(2,1,1) cases together. Thus we have U=(u1,u2)U=(u_{1},u_{2}), where the variables appearing in u1u_{1} are independent of those in u2u_{2} and each has two columns, then W=(u2u1)W=\begin{pmatrix}-u_{2}\\ u_{1}\end{pmatrix}. By our analysis above, in order to have a two dimensional linear annihilator, either u1=L1=μIdu_{1}=L_{1}=\mu\operatorname{Id} and u2u_{2} can be anything, or u1u_{1} and u2u_{2} are J2,0J_{2,0}’s. Note that in either case, u1,u2u_{1},u_{2} must commute. We have the following possibilities:

u1=L1u_{1}=L_{1} and u2u_{2} anything, i.e., one of L1,J2,0,J2,λ,L1𝕥L1𝕥,L1𝕥J1,λ,J1,λJ1,λ,J1,λJ1,0L1𝕥J1,0,J1,0J1,0L_{1},J_{2,0},J_{2,\lambda},L_{1}^{\mathbb{t}}L_{1}^{\mathbb{t}},L_{1}^{\mathbb{t}}J_{1,\lambda},J_{1,\lambda}J_{1,\lambda^{\prime}},J_{1,\lambda}J_{1,0}L_{1}^{\mathbb{t}}J_{1,0},J_{1,0}J_{1,0} where λ,λ0\lambda,\lambda^{\prime}\neq 0, or

u1=J2,0u_{1}=J_{2,0} and u2=J2,0u_{2}=J_{2,0}.

Note the only cases with u2u_{2} non-invertible are u2=J1,0J1,0u_{2}=J_{1,0}J_{1,0} and u2=J1,λJ1,λu_{2}=J_{1,\lambda}J_{1,\lambda} (same λ\lambda in both).

Write 𝕩=(x1x2x3x4){\mathbb{x}}=\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix} with each xjx_{j} 2×22\times 2, then the U𝕩W=0U{\mathbb{x}}W=0 equation gives the equation of 2×22\times 2 matrices with cubic entries:

u1x1u2u2x3u2+u1x2u1+u2x4u1=0-u_{1}x_{1}u_{2}-u_{2}x_{3}u_{2}+u_{1}x_{2}u_{1}+u_{2}x_{4}u_{1}=0

Write xj=yj+xj(u1)+xj(u2)x_{j}=y_{j}+x_{j}(u_{1})+x_{j}(u_{2}) where the variables in yjy_{j} are independent of those in u1,u2u_{1},u_{2} and xj(ui)x_{j}(u_{i}) has entries linear in the entries of uiu_{i}. We immediately obtain y2=x2(u1)=0y_{2}=x_{2}(u_{1})=0 and if u2J1,0J1,0u_{2}\neq J_{1,0}J_{1,0}, y3=x3(u2)=0y_{3}=x_{3}(u_{2})=0.

Note we are free to modify x1x_{1} (resp. x3x_{3}) by a multiple of u1u_{1} as long as we modify x2x_{2} (resp. x4)x_{4}), by the same multiple of u2u_{2}, and x1x_{1} (resp. x2x_{2}) by a multiple of u2-u_{2} as long as we modify x3x_{3} (resp. x4x_{4}) by the same multiple of u1u_{1}.

Consider the case u1=u2=J2,0u_{1}=u_{2}=J_{2,0}. Note that the centralizer (among matrices of linear forms) of a J2,0J_{2,0} is spanned by tIdt\operatorname{Id} and J2,0J_{2,0}. Separating the equations by the variables involved, we obtain y1=y4=tId+J2,0y_{1}=y_{4}=t\operatorname{Id}+J_{2,0}, and we may absorb the tIdt\operatorname{Id} into the J2,0J_{2,0}. Taking into account the other homogeneities and our allowed modifications, we obtain Case (IV).

Next consider the case u1=L1u_{1}=L_{1} and u2u_{2} has centralizer tIdt\operatorname{Id}, which occurs when u2u_{2} is any of J2,λ,L1𝕥L1𝕥,L1𝕥J1,λ,L1𝕥J1,0J_{2,\lambda},L_{1}^{\mathbb{t}}L_{1}^{\mathbb{t}},L_{1}^{\mathbb{t}}J_{1,\lambda},L_{1}^{\mathbb{t}}J_{1,0}. In all these cases we are reduced to 𝕩=tId4{\mathbb{x}}=t\operatorname{Id}_{4}, but the resulting spaces are all specializations of the case u2=L1𝕥L1𝕥u_{2}=L_{1}^{\mathbb{t}}L_{1}^{\mathbb{t}}, and thus not basic unless u2=L1𝕥L1𝕥u_{2}=L_{1}^{\mathbb{t}}L_{1}^{\mathbb{t}}, which is Case (III).

Case (u1,u2)=(L1,L1)(u_{1},u_{2})=(L_{1},L_{1}): the solutions to the homogeneous parts of the equations, after admissible modifications, give y1=y4y_{1}=y_{4}, and all other terms in 𝕩{\mathbb{x}} are zero. The resulting space is a permuted version of Case (III).

Case u1=L1u_{1}=L_{1}, u2=J2,0u_{2}=J_{2,0}. This case gives a specialization of Case (IV).

Case u1=L1u_{1}=L_{1}, u2=J1,0J1,0u_{2}=J_{1,0}J_{1,0}. The U𝕩W=0U{\mathbb{x}}W=0 equation implies x22x^{2}_{2} is the only nonzero entry in the second row, so this case is not primitive.

The case (1,1,1,1)(1,1,1,1) is easily ruled out.

6.6. (6,5)(6,5) case

The same argument as in the (5,7)(5,7) case shows that UU cannot have any columns equal to zero. Thanks to Proposition 5.6, the cases that might give basic spaces are when UU has a one-dimensional linear annihilator and atR=2at_{R}=2, so 𝕩W{\mathbb{x}}W must not be a linear form times WW, and thus UU has a primitive degree two annihilator. By the discussion in §6.4, there are four cases with a one-dimensional linear annihilator: (L2,J1,0)(L_{2},J_{1,0}), (J3,0,J1,0)(J_{3,0},J_{1,0}), (J2,0,J1,0,L1𝕥)(J_{2,0},J_{1,0},L_{1}^{\mathbb{t}}), and (J1,0,J1,0,L1𝕥,L1𝕥)(J_{1,0},J_{1,0},L_{1}^{\mathbb{t}},L_{1}^{\mathbb{t}}) which we may respectively write as

(u1u20u30u1u20),(u1u2u3u40u1u20),(u1u2u3u40u10u5),(u1u2u3u400u5u6).\begin{pmatrix}u_{1}&u_{2}&0&u_{3}\\ 0&u_{1}&u_{2}&0\end{pmatrix},\ \begin{pmatrix}u_{1}&u_{2}&u_{3}&u_{4}\\ 0&u_{1}&u_{2}&0\end{pmatrix},\ \begin{pmatrix}u_{1}&u_{2}&u_{3}&u_{4}\\ 0&u_{1}&0&u_{5}\end{pmatrix},\ \begin{pmatrix}u_{1}&u_{2}&u_{3}&u_{4}\\ 0&0&u_{5}&u_{6}\end{pmatrix}.

Note that the first is a degeneration of the second, and after permuting columns, the second is a degeneration of the third, and the third a degeneration of the fourth. To make the degeneration transparent, we rename the variables and write the spaces as follows

(u1u2u3000u1u3),(u1u2u3u400u1u3),(u1u2u3u400u1u6),(u1u2u3u400u5u6).\begin{pmatrix}u_{1}&u_{2}&u_{3}&0\\ 0&0&u_{1}&u_{3}\end{pmatrix},\ \begin{pmatrix}u_{1}&u_{2}&u_{3}&u_{4}\\ 0&0&u_{1}&u_{3}\end{pmatrix},\ \begin{pmatrix}u_{1}&u_{2}&u_{3}&u_{4}\\ 0&0&u_{1}&u_{6}\end{pmatrix},\ \begin{pmatrix}u_{1}&u_{2}&u_{3}&u_{4}\\ 0&0&u_{5}&u_{6}\end{pmatrix}.

In all the cases, W=(u2,u1,0,0)𝕥W=(-u_{2},u_{1},0,0)^{\mathbb{t}}. We computed by Macaulay2 that primitive degree two annihilator of UU is generated by (0,u4u5+u3u6,u2u6,u2u5)𝕥(0,-u_{4}u_{5}+u_{3}u_{6},-u_{2}u_{6},u_{2}u_{5})^{\mathbb{t}} and (u4u5+u3u6,0,u1u6,u1u5)𝕥(-u_{4}u_{5}+u_{3}u_{6},0,-u_{1}u_{6},u_{1}u_{5})^{\mathbb{t}}, where the effect of degeneration does not change the space of primitive degree two annihilators. By analyzing the first entry of 𝕩W{\mathbb{x}}W, we see that 𝕩W{\mathbb{x}}W has to be a linear form times WW since it cannot involve any primitive degree two annihilators, which is a contradiction.

6.7. Good Chern classes

By construction the Atkinson numbers in our examples are all 22, here we see the Chern classes are as well. We work in a slightly more general setting where u1,u2,x1u_{1},u_{2},x_{1} are commuting k×kk\times k matrices, i.e., the general setting of Proposition 2.3 applied to 3Hom(3,Λ23)\mathbb{C}^{3}\subset\operatorname{Hom}(\mathbb{C}^{3},\Lambda^{2}\mathbb{C}^{3}). Our space is

(x1u2x1u1u1u2)\begin{pmatrix}x_{1}&&-u_{2}\\ &x_{1}&u_{1}\\ u_{1}&u_{2}&\end{pmatrix}

Then the left kernel is the k×3kk\times 3k matrix of linear forms (u1,u2,x1)(-u_{1},-u_{2},x_{1}), and the right kernel is the 3k×k3k\times k matrix (u2u1x1)\begin{pmatrix}u_{2}\\ -u_{1}\\ x_{1}\end{pmatrix}. Since the blocks u1,u2,x1u_{1},u_{2},x_{1} are each in different sets of variables, the matrix of size kk minors has no common factor and we conclude c1()=c1()=kc_{1}({\mathcal{E}})=c_{1}({\mathcal{F}})=k.

6.8. Proof that our two new examples are basic

Proof.

By the Buchsbaum-Eisenbud criterion for exactness [5], it is easy to see that the following complexes are exact as well as their duals:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}R2\textstyle{R^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d3\scriptstyle{d_{3}}R6\textstyle{R^{6}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Case(III)\scriptstyle{Case(III)}R6\textstyle{R^{6}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1\scriptstyle{d_{1}}R2.\textstyle{R^{2}.}
0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}R2\textstyle{R^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d3\scriptstyle{d_{3}^{{}^{\prime}}}R6\textstyle{R^{6}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Case(IV)\scriptstyle{Case(IV)}R6\textstyle{R^{6}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1\scriptstyle{d_{1}^{{}^{\prime}}}R2.\textstyle{R^{2}.}

where R=[a1,,a6]R=\mathbb{C}[a_{1},\ldots,a_{6}],

d3=(a3a5a4a6a200a2a100a1),d1=(a20a3a5a100a2a4a60a1),d3=(a5a60a5a3a40a3a1a20a1),andd1=(a3a4a5a6a1a20a30a50a1).d_{3}=\left(\begin{smallmatrix}a_{3}&a_{5}\\ a_{4}&a_{6}\\ -a_{2}&0\\ 0&a_{2}\\ a_{1}&0\\ 0&a_{1}\end{smallmatrix}\right),\ d_{1}=\left(\begin{smallmatrix}-a_{2}&0&-a_{3}&-a_{5}&a_{1}&0\\ 0&-a_{2}&-a_{4}&-a_{6}&0&a_{1}\end{smallmatrix}\right),\ d_{3}^{{}^{\prime}}=\left(\begin{smallmatrix}a_{5}&a_{6}\\ 0&a_{5}\\ -a_{3}&-a_{4}\\ 0&-a_{3}\\ a_{1}&a_{2}\\ 0&a_{1}\end{smallmatrix}\right),{\rm\ and\ }d_{1}^{{}^{\prime}}=\left(\begin{smallmatrix}-a_{3}&-a_{4}&-a_{5}&-a_{6}&a_{1}&a_{2}\\ 0&-a_{3}&0&-a_{5}&0&a_{1}\end{smallmatrix}\right).

This proves Case (III) and Case (IV) are unexpandable.

The spaces Case (III) and Case (IV) are strongly indecomposable if and only if the image sheaves of d1,d1d_{1},d_{1}^{{}^{\prime}} as well as the image sheaf of d3𝕥,(d3)𝕥d_{3}^{\mathbb{t}},(d_{3}^{{}^{\prime}})^{\mathbb{t}} are indecomposable. In the case corresponding to Case (III), we compute the ideal generated by maximal minors of d3d_{3} as well as d1d_{1}. They each have 1212 minimal generators. However, if their image sheaf were decomposable, the number of minimal generators would be at most (52)=10\binom{5}{2}=10. To see this, one obtains the most generators when the last column of d3𝕥d_{3}^{\mathbb{t}} or d1d_{1} is zero. In the case corresponding to Case (IV), note that the entries of the first row vector of d1d_{1}^{{}^{\prime}} as well as the entries of the second column vector of d3d_{3}^{{}^{\prime}} are algebraically independent (even after any possible row operations). Hence up to row/column operations and GL(A)\mathrm{GL}(A) actions, we cannot have a zero entry in the first row of d1d_{1}^{{}^{\prime}} (respectively, the second column of d3d_{3}^{{}^{\prime}}). Thus the image sheaves of d1d_{1}^{{}^{\prime}} and d3d_{3}^{{}^{\prime}} are indecomposable.

Finally, to show Case (III) and Case (IV) are unliftable, we used code in Macaulay2 implementing Proposition 4.3. ∎

7. Additional Examples

We first give two corank two examples generalizing the new rank four cases: If there are exactly two L1L_{1} blocks and at most one Jor1,0Jor_{1,0} block, then there is a 2𝐛62\mathbf{b}-6 dimensional kernel, and the resulting space of bounded rank 𝐛2\mathbf{b}-2 size 𝐛×(2𝐛6)\mathbf{b}\times(2\mathbf{b}-6) matrices is a specialization of the case with 𝐛4\mathbf{b}-4 L1𝕥L_{1}^{\mathbb{t}}’s, so it is sufficient consider that space. That is, set p=𝐛3p=\mathbf{b}-3,the space is

(13) (a10a20a3a5a2p10a10a2a4a6a2p)\begin{pmatrix}a_{1}&0&a_{2}&0&a_{3}&a_{5}&\cdots&a_{2p-1}\\ 0&a_{1}&0&a_{2}&a_{4}&a_{6}&\cdots&a_{2p}\end{pmatrix}

and after permuting rows of the kernel matrix to make it in Atkinson normal form one obtains

(14) (a1a3a5a2p1a1a4a6a2pa2a1a2a1a1a2a20a3a5a2p1000a2a4a6a2p00)\begin{pmatrix}a_{1}&&&&&&-a_{3}&-a_{5}&\cdots&-a_{2p-1}\\ &a_{1}&&&&&-a_{4}&-a_{6}&\cdots&-a_{2p}\\ &&\ddots&&&&a_{2}&&&\\ &&&a_{1}&&&&a_{2}&&\\ &&&&a_{1}&&&&\ddots&\\ &&&&&a_{1}&&&&a_{2}\\ -a_{2}&0&-a_{3}&-a_{5}&\cdots&-a_{2p-1}&0&\cdots&&0\\ 0&-a_{2}&-a_{4}&-a_{6}&\cdots&-a_{2p}&0&\cdots&&0\end{pmatrix}

This is a 2(𝐛3)2(\mathbf{b}-3)-dimensional space of 𝐛×2𝐛6\mathbf{b}\times 2\mathbf{b}-6 matrices of bounded rank 𝐛2\mathbf{b}-2 generalizing Case (III).

The example Case (IV) generalizes to a 3q3q-dimensional space of [2q+(q2)]×[2q+2][2q+\binom{q}{2}]\times[2q+2] matrices of bounded rank 2q2q.

More generally, going beyond corank two using the same blow-up, with XX a k×kk\times k matrix of linear forms

(a1IdkXa1Idka2Idka2IdkX0)\begin{pmatrix}a_{1}\operatorname{Id}_{k}&&-X\\ &a_{1}\operatorname{Id}_{k}&a_{2}\operatorname{Id}_{k}\\ a_{2}\operatorname{Id}_{k}&X&0\end{pmatrix}

gives a k2+2k^{2}+2 dimensional space of 3k×3k3k\times 3k matrices of bounded rank 2k2k. L. Manivel points out that this suggests (after permuting the blocks):

(a1Idka2Idka3IdkX(a2a3)IdkX(a3a1)IdkX(a1a2)Idk)\begin{pmatrix}a_{1}\operatorname{Id}_{k}&a_{2}\operatorname{Id}_{k}&a_{3}\operatorname{Id}_{k}\\ X&&&(a_{2}-a_{3})\operatorname{Id}_{k}\\ &X&&(a_{3}-a_{1})\operatorname{Id}_{k}\\ &&X&(a_{1}-a_{2})\operatorname{Id}_{k}\end{pmatrix}

and more generally pp blocks with XX and (a1,,ak)(a_{1},\dotsc,a_{k}), (b1,,bk)(b_{1},\dotsc,b_{k}) linear forms such that

(a1,,ak)(b1bk)=0.(a_{1},\dotsc,a_{k})\begin{pmatrix}b_{1}\\ \vdots\\ b_{k}\end{pmatrix}=0.

There is a unique up to isomorphism concise tensor in 333\mathbb{C}^{3}{\mathord{\otimes}}\mathbb{C}^{3}{\mathord{\otimes}}\mathbb{C}^{3} that is both of minimal border rank and gives rise to a space of bounded rank [7], the space is

(a1a2a3a1a3)\begin{pmatrix}a_{1}&a_{2}&a_{3}\\ &a_{1}&\\ &a_{3}&\end{pmatrix}

(the tensor is 1B1_{B} and 1C1_{C}-generic, and 1A1_{A}-degenerate). Applying the construction of Proposition 2.3 with the other tensor Tskewcw,2T_{skewcw,2}, yields a bounded rank 66 space in 99\mathbb{C}^{9}{\mathord{\otimes}}\mathbb{C}^{9}. This space is compression, but it may be useful for Strassen’s laser method. It also shows that even if the space of k×kk\times k matrices is of bounded rank less than kk, one can still obtain a bounded rank krkr space with the construction.

References

  • [1] Andris Ambainis, Yuval Filmus, and François Le Gall, Fast matrix multiplication: limitations of the Coppersmith-Winograd method (extended abstract), STOC’15—Proceedings of the 2015 ACM Symposium on Theory of Computing, ACM, New York, 2015, pp. 585–593. MR 3388238
  • [2] M. D. Atkinson, Primitive spaces of matrices of bounded rank. II, J. Austral. Math. Soc. Ser. A 34 (1983), no. 3, 306–315. MR MR695915 (84h:15017)
  • [3] M. D. Atkinson and S. Lloyd, Large spaces of matrices of bounded rank, Quart. J. Math. Oxford Ser. (2) 31 (1980), no. 123, 253–262. MR 587090
  • [4] Markus Bläser and Vladimir Lysikov, Slice Rank of Block Tensors and Irreversibility of Structure Tensors of Algebras, 45th International Symposium on Mathematical Foundations of Computer Science (MFCS 2020) (Dagstuhl, Germany) (Javier Esparza and Daniel Kráľ, eds.), Leibniz International Proceedings in Informatics (LIPIcs), vol. 170, Schloss Dagstuhl–Leibniz-Zentrum für Informatik, 2020, pp. 17:1–17:15.
  • [5] David A. Buchsbaum and David Eisenbud, What makes a complex exact?, J. Algebra 25 (1973), 259–268. MR 314819
  • [6] by same author, Algebra structures for finite free resolutions, and some structure theorems for ideals of codimension 33, Amer. J. Math. 99 (1977), no. 3, 447–485. MR 453723
  • [7] Jarosław Buczyński and J. M. Landsberg, On the third secant variety, J. Algebraic Combin. 40 (2014), no. 2, 475–502. MR 3239293
  • [8] Austin Conner, Fulvio Gesmundo, Joseph M. Landsberg, Emanuele Ventura, and Yao Wang, Towards a geometric approach to Strassen’s asymptotic rank conjecture, Collect. Math. 72 (2021), no. 1, 63–86. MR 4204580
  • [9] Austin Conner, Alicia Harper, and J.M. Landsberg, New lower bounds for matrix multiplication and det3\operatorname{det}_{3}, Forum of Mathematics, Pi 11 (2023), e17.
  • [10] Jan Draisma, Small maximal spaces of non-invertible matrices, Bull. London Math. Soc. 38 (2006), no. 5, 764–776. MR 2268360 (2007j:17010)
  • [11] David Eisenbud and Joe Harris, On varieties of minimal degree (a centennial account), Algebraic geometry, Bowdoin, 1985 (Brunswick, Maine, 1985), Proc. Sympos. Pure Math., vol. 46, Amer. Math. Soc., Providence, RI, 1987, pp. 3–13. MR 927946
  • [12] by same author, Vector spaces of matrices of low rank, Adv. in Math. 70 (1988), no. 2, 135–155. MR MR954659 (89j:14010)
  • [13] H. Flanders, On spaces of linear transformations with bounded rank, J. London Math. Soc. 37 (1962), 10–16. MR 136618
  • [14] Runshi Geng and Joseph M. Landsberg, On the geometry of geometric rank, Algebra Number Theory 16 (2022), no. 5, 1141–1160. MR 4471039
  • [15] Joachim Jelisiejew, J. M. Landsberg, and Arpan Pal, Concise tensors of minimal border rank, Math. Ann. (2022).
  • [16] Swastik Kopparty, Guy Moshkovitz, and Jeroen Zuiddam, Geometric Rank of Tensors and Subrank of Matrix Multiplication, 35th Computational Complexity Conference (CCC 2020) (Dagstuhl, Germany) (Shubhangi Saraf, ed.), Leibniz International Proceedings in Informatics (LIPIcs), vol. 169, Schloss Dagstuhl–Leibniz-Zentrum für Informatik, 2020, pp. 35:1–35:21.
  • [17] J. M. Landsberg, Geometry and complexity theory, Cambridge Studies in Advanced Mathematics, vol. 169, Cambridge University Press, Cambridge, 2017. MR 3729273
  • [18] J. M. Landsberg and L. Manivel, The sextonions and E712E_{7\frac{1}{2}}, Adv. Math. 201 (2006), no. 1, 143–179. MR 2204753 (2006k:17017)
  • [19] J. M. Landsberg and L. Manivel, Equivariant spaces of matrices of constant rank, 2022.
  • [20] J. M. Landsberg and Mateusz Michałek, On the geometry of border rank decompositions for matrix multiplication and other tensors with symmetry, SIAM J. Appl. Algebra Geom. 1 (2017), no. 1, 2–19. MR 3633766
  • [21] J. M. Landsberg and Mateusz Michałek, Towards finding hay in a haystack: explicit tensors of border rank greater than 2.02m2.02m in cmcmcmc^{m}\otimes c^{m}\otimes c^{m}, 2019, arXiv1912.11927.
  • [22] Joseph M. Landsberg and Mateusz Michałek, A 2n2log2(n)12n^{2}-\log_{2}(n)-1 lower bound for the border rank of matrix multiplication, Int. Math. Res. Not. IMRN (2018), no. 15, 4722–4733. MR 3842382
  • [23] L. Manivel and R. M. Roig, Matrices of linear forms of constant rank from vector bundles on projective space, 2023.
  • [24] V. Strassen, Relative bilinear complexity and matrix multiplication, J. Reine Angew. Math. 375/376 (1987), 406–443. MR MR882307 (88h:11026)
  • [25] by same author, Algebra and complexity, First European Congress of Mathematics, Vol. II (Paris, 1992), Progr. Math., vol. 120, Birkhäuser, Basel, 1994, pp. 429–446. MR 1341854
  • [26] John Sylvester, On the dimension of spaces of linear transformations satisfying rank conditions, Linear Algebra Appl. 78 (1986), 1–10. MR 840165
  • [27] Terrance Tao, A symmetric formulation of the croot-lev-pach-ellenberg-gijswijt capset bound, https://terrytao.wordpress.com/2016/05/18/a-symmetric-formulation-of-the-croot-lev-pach-ellenberg-gijswijt-capset-bound.
  • [28] R. Westwick, Spaces of matrices of fixed rank. II, Linear Algebra Appl. 235 (1996), 163–169. MR 1374258 (97g:15001)