math0607045/lfd.tex
1: \documentclass[anglais]{cedram-aif}
2: 
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: % version
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: 
7: \usepackage{version}
8: \excludeversion{annalif}
9: \includeversion{arxiv}
10: 
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
12: % AIF packages
13: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
14: 
15: \usepackage[english]{babel}
16: \usepackage[latin1]{inputenc}
17: \usepackage[T1]{fontenc}
18: 
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20: % packages
21: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
22: 
23: \usepackage{amssymb,amsmath,amsthm}
24: \usepackage{epic,eepic,longtable}
25: \usepackage{dsfont,mathrsfs}
26: \usepackage[neveradjust]{paralist}
27: \usepackage{url}
28: \input{xypic}
29: \xyoption{all}
30: 
31: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32: % abbreviations
33: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
34: 
35: \newcommand{\bu}{\bullet}
36: \newcommand{\ssm}{\smallsetminus}
37: \newcommand{\beq}{\begin{equation}}
38: \newcommand{\eeq}{\end{equation}}
39: \newcommand{\eps}{\varepsilon}
40: \newcommand{\ideal}[1]{{\langle#1\rangle}}
41: \newcommand{\into}{\hookrightarrow}
42: \newcommand{\onto}{\twoheadrightarrow}
43: \newcommand{\w}{\wedge}
44: \newcommand{\p}{\partial}
45: \newcommand{\xymat}{\SelectTips{cm}{}\xymatrix}
46: 
47: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48: % functions
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: 
51: \newcommand{\wh}[1]{{\widehat{#1}}}
52: \newcommand{\del}[2]{\frac{\p}{\p x_{#1,#2}}}
53: 
54: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
55: % greek
56: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57: 
58: \newcommand{\ve}{\varepsilon}
59: \newcommand{\vp}{\varphi}
60: \newcommand{\Ga}{\Gamma}
61: \newcommand{\Om}{\Omega}
62: \newcommand{\om}{\omega}
63: 
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: % numbers
66: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67: 
68: \renewcommand{\mathbb}{\mathds}
69: \newcommand{\QQ}{\mathbb{Q}}
70: \newcommand{\RR}{\mathbb{R}}
71: \newcommand{\CC}{\mathbb{C}}
72: \newcommand{\ZZ}{\mathbb{Z}}
73: \newcommand{\PP}{\mathbb{P}}
74: \newcommand{\NN}{\mathbb{N}}
75: 
76: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
77: % sheaves
78: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
79: 
80: \renewcommand{\mathcal}{\mathscr}
81: \newcommand{\A}{\mathcal{A}}
82: \newcommand{\C}{\mathcal{C}}
83: \newcommand{\D}{\mathcal{D}}
84: \renewcommand{\O}{\mathcal{O}}
85: \renewcommand{\P}{\mathcal{P}}
86: \newcommand{\R}{\mathcal{R}}
87: 
88: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89: % Lie algebra
90: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91: 
92: \newcommand{\gl}{\mathfrak{gl}}
93: \newcommand{\pgl}{\mathfrak{pgl}}
94: \renewcommand{\sl}{\mathfrak{sl}}
95: 
96: \renewcommand{\aa}{\mathfrak{a}}
97: \newcommand{\bb}{\mathfrak{b}}
98: \newcommand{\dd}{\mathfrak{d}}
99: \renewcommand{\gg}{\mathfrak{g}}
100: \newcommand{\hh}{\mathfrak{h}}
101: \newcommand{\kk}{\mathfrak{k}}
102: \renewcommand{\L}{\mathrm{L}}
103: \newcommand{\mm}{\mathfrak{m}}
104: \renewcommand{\ss}{\mathfrak{s}}
105: \newcommand{\ttt}{\mathfrak{t}}
106: \newcommand{\zz}{\mathfrak{z}}
107: 
108: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
109: % quiver
110: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
111: 
112: \renewcommand{\d}{\mathbf{d}}
113: \newcommand{\ii}{\mathrm{i}}
114: 
115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
116: % operators
117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
118: 
119: \DeclareMathOperator{\ad}{ad}
120: \DeclareMathOperator{\Aut}{Aut}
121: \DeclareMathOperator{\cod}{codim}
122: \DeclareMathOperator{\depth}{depth}
123: \DeclareMathOperator{\Der}{Der}
124: \DeclareMathOperator{\diag}{diag}
125: \DeclareMathOperator{\End}{End}
126: \DeclareMathOperator{\Ext}{Ext}
127: \DeclareMathOperator{\Gl}{Gl}
128: \DeclareMathOperator{\gr}{gr}
129: \DeclareMathOperator{\Hom}{Hom}
130: \DeclareMathOperator{\im}{im}
131: \DeclareMathOperator{\Mat}{Mat}
132: \DeclareMathOperator{\pd}{pd}
133: \DeclareMathOperator{\Rep}{Rep}
134: \DeclareMathOperator{\sgn}{sign}
135: \DeclareMathOperator{\tr}{trace}
136: \DeclareMathOperator{\Sing}{Sing}
137: \DeclareMathOperator{\Sl}{Sl}
138: \DeclareMathOperator{\Spec}{Spec}
139: \DeclareMathOperator{\supp}{supp}
140: \DeclareMathOperator{\Sym}{Sym}
141: \DeclareMathOperator{\vol}{vol}
142: \DeclareMathOperator{\wt}{wt}
143: \DeclareMathOperator{\Z}{Z}
144: 
145: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
146: % Title, Authors, Classification
147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
148: 
149: \title[Linear free divisors]{Linear free divisors and the\\global logarithmic comparison theorem}
150: \alttitle{Diviseurs lin\'eairement libres\\ et le th\'eor\`eme de\\ comparaison logarithmique global}
151: 
152: \author{\firstname{Michel} \lastname{Granger}}
153: \address{
154: Departement de Math\'ematiques\\
155: Universit\'e d'Angers\\
156: 2 Bd Lavoisier\\
157: 49045 Angers\\
158: France
159: }
160: \email{granger@univ-angers.fr}
161: \thanks{}
162: 
163: \author{\firstname{David} \lastname{Mond}}
164: \address{
165: Mathematics Institute\\
166: University of Warwick\\
167: Coventry CV47AL\\
168: England
169: }
170: \email{D.M.Q.Mond@warwick.ac.uk}
171: \thanks{DM is grateful to Ignacio de Gregorio for helpful conversations on the topics treated here.}
172: 
173: \author{\firstname{Alicia} \lastname{Nieto-Reyes}}
174: \address{
175: Departamento de Matematicas,\\
176: Estadistica y Computacion\\
177: Universidad de Cantabria\\
178: Spain
179: }
180: \email{alicia.nieto@unican.es}
181: \thanks{}
182: 
183: \author{\firstname{Mathias} \lastname{Schulze}}
184: \address{
185: Department of Mathematics\\
186: Oklahoma State University\\
187: Stillwater, OK 74078\\
188: United States}
189: \email{mschulze@math.okstate.edu}
190: \thanks{MS gratefully acknowledges financial support from EGIDE and the Humboldt Foundation.}
191: 
192: \thanks{We are grateful to the referee for a very careful reading and many valuable suggestions.}
193: 
194: \keywords{free divisor, prehomogeneous vector space, de~Rham cohomology, logarithmic comparison theorem, Lie algebra cohomology, quiver representation}
195: \altkeywords{diviseur lin\'eairement libre, espace vectorielle pr\'ehomog\`ene, cohomologie de de~Rham, th\'eor\`eme de comparaison logarithmique, cohomologie des alg\`ebres de Lie, repr\'esentation des quivers}
196: 
197: \subjclass{32S20, 14F40, 20G10, 17B66}
198: % 32S20 Global theory of singularities; cohomological properties
199: % 14F40 de Rham cohomology
200: % 20G10 Linear algebraic groups (classical groups) / Cohomology theory
201: % 17B66 Lie algebras of vector fields and related (super) algebras
202: 
203: \date{January 16, 2008}
204: 
205: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
206: \begin{document}
207: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
208: 
209: \begin{abstract}
210: A complex hypersurface $D$ in $\CC^n$ is a \emph{linear free divisor} (LFD) if its module of logarithmic vector fields has a global basis of linear vector fields. We classify all LFDs for $n$ at most $4$. 
211: 
212: By analogy with Grothendieck's comparison theorem, we say that the \emph{global logarithmic comparison theorem} (GLCT) holds for $D$ if the complex of global logarithmic differential forms computes the complex cohomology of $\CC^n\setminus D$.
213: We develop a general criterion for the GLCT for LFDs and prove that it is fulfilled whenever the Lie algebra of linear logarithmic vector fields is reductive.
214: For $n$ at most $4$, we show that the GLCT holds for all LFDs. 
215: 
216: We show that LFDs arising naturally as discriminants in quiver representation spaces (of real Schur roots) fulfill the GLCT. As a by-product we obtain a topological proof of a theorem of V.~Kac on the number of irreducible components of such discriminants.
217: \end{abstract}
218: 
219: \begin{altabstract}
220: Une hypersurface complexe de $\CC^n$ est appel\'ee un {\it diviseur lin\'eairement libre} (ou DLL) si son module de champs de vecteur logarithmiques a une base globale form\'ee de champs de vecteurs lin\'eaires. 
221: Nous classifions tous les DLL pour $n$ au plus \'egal a $4$.
222: 
223: Par analogie avec le th\'eor\`eme de comparaison de Grothendieck, on dit que le {\it th\'eor\`eme de comparaison logarithmique global} (ou TCLG) est vrai pour $D$ si le complexe des formes diff\'erentielles logarithmiques globales permet de calculer la 
224: cohomologie de $\CC^n\setminus D$ \`a coefficients complexes. 
225: Nous mettons en \'evidence un crit\`ere g\'en\'eral pour qu'un DLL ait la propri\'et\'e  TCLG, et nous d\'emontrons que ce crit\`ere s'applique lorsque l'alg\`ebre de Lie des champs de vecteurs logarithmiques lin\'eaires est r\'eductive. 
226: Pour $n$ inf\'erieur ou \'egal \`a $4$, nous montrons que le TCLG est vrai pour tous les DLL.
227: 
228: Nous montrons que les DLL qui apparaissent naturellement comme discriminants dans les espaces de repr\'esentations de carquois pour des racines de Schur r\'eelles satisfont au TCLG. 
229: Comme corollaire nous obtenons une d\'emonstration topologique d'un r\'esultat de V. Kac sur le nombre de composantes irr\'eductibles de tels discriminants.
230: \end{altabstract}
231: 
232: \maketitle
233: 
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: \section{Introduction}
236: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
237: 
238: We denote by $\O=\O_{\CC^n}$ the sheaf of holomorphic functions on $\CC^n$, by $\mm_p\subseteq\O_p$ the maximal ideal at $p\in\CC^n$, by $\Der=\Der_{\CC}(\O)$ the sheaf of $\CC$-linear derivations of $\O$ (or \emph{holomorphic vector fields}) on $\CC^n$, and by $\Om^\bullet=\Om^\bullet_{\CC^n}$ the complex of sheaves of holomorphic differential forms.
239: We shall frequently use a local or global coordinate system $x=x_1,\dots,x_n$ on $\CC^n$ and then denote by $\p=\p_1,\dots,\p_n$ the corresponding operators of partial derivatives $\p_i=\frac{\p}{\p x_i}$, $i=1,\dots,n$.
240: Note that $\Der=\bigoplus\O\cdot\p_i$ is a free $\O$-module of rank $n$.
241: 
242: Let $D\subseteq\CC^n$ be a reduced divisor.
243: K.~Saito \cite{Sai80} associated to $D$ the (coherent) sheaf of \emph{logarithmic vector fields} $\Der(-\log D)\subseteq\Der$ and the complex of (coherent) sheaves $\Om^\bu(\log D)\subseteq\Om^\bu(*D)$ of \emph{logarithmic differential forms} along $D$.
244: For a (local or global) defining equation $\Delta\in\O$ of the germ $D$, $\delta\in\Der$ is in $\Der(-\log D)$ if $\delta(\Delta)\in\O\cdot\Delta$, and $\om\in\Om^\bullet[\Delta^{-1}]$ is in $\Om^\bu(\log D)$ if $\Delta\cdot\om,\Delta\cdot d\om\in\Om^\bullet$.
245: Note that $\Der(-\log D)$ contains the \emph{annihilator} $\Der(-\log\Delta)$ of $\Delta$ defined by the condition $\delta(\Delta)=0$.
246: Saito showed that $\Der(-\log D)$ and $\Om^1(\log D)$ are reflexive and mutually dual and introduced the following important class of divisors.
247: 
248: \begin{defi}
249: A divisor $D$ is called \emph{free} if $\Der(-\log D)$, or equivalently $\Om^1(\log D)$, is a locally free $\O$-module, necessarily of rank $n$. 
250: \end{defi}
251: 
252: We will be concerned in this article with the following subclass of divisors.
253: 
254: \begin{defi}\label{2}
255: A free divisor $D$ is called \emph{linear} if $\Ga(\CC^n,\Der(-\log D))$ admits a basis $\delta_1,\dots,\delta_n$ such that each $\delta_i$ has linear coefficients with respect to the $\O$-basis $\p_1,\dots,\p_n$ of $\Der$ or equivalently each $\delta_i$ is homogeneous of degree zero with respect to the standard degree defined by $\deg x_i=1=-\deg\p_i$ on the variables and generators of $\Der$.
256: \end{defi}
257: 
258: Saito's criterion \cite[Thm.~1.8.(ii)]{Sai80} implies the following fundamental observation.
259: 
260: \begin{lemm}\label{28}
261: If $\delta_1,\dots,\delta_n$ is a basis of $\Ga(\CC^n,\Der(-\log D))$ for a linear free divisor $D$, then the homogeneous polynomial $\Delta=\det((\delta_i(x_j))_{i,j})\in\CC[x]$ of degree $n$ is a global defining equation for $D$.
262: \end{lemm}
263: 
264: Note that because $\Der(-\log D)$ can have no members of negative degree, $D$ cannot be isomorphic to the product of $\CC$ with a lower dimensional divisor. 
265: It turns out that linear free divisors are relatively abundant; the authors believe that in the current paper and in \cite{BM06}, recipes are given which allow the straightforward construction of more free divisors than have been described in the sum of all previous papers. 
266: 
267: \begin{exems}\label{1}\
268: \begin{asparaenum}
269: \item The normal crossing divisor $D=\{x_1\cdots x_n=0\}\subseteq\CC^n$ is a linear free divisor where
270: \[
271: x_1\p_1,\dots, x_n\p_n
272: \]
273: is a basis of $\Der(-\log D)$.
274: Up to isomorphism it is the only example among hyperplane arrangements, cf.~\cite[Ch.~4]{OT92}.
275: 
276: \item\label{1b} In the space $B_{2,3}$ of binary cubics, the discriminant $D$, which consists of binary cubics having a repeated root, is a linear free divisor. 
277: For $f(u,v)=xu^3+yu^2v+zuv^2+wv^3$ has a repeated root if and only if its Jacobian ideal does not contain any power of the maximal ideal $(u,v)$, and this in turn holds if and only if the four cubics
278: \[
279: u\p_uf,v\p_uf,u\p_vf,v\p_vf
280: \]
281: are linearly dependent. 
282: Writing the coefficients of these four cubics as the columns of the $4\times 4$ matrix
283: \[
284: A:=
285: \begin{pmatrix}
286: 3x&0&y&0\\
287: 2y&3x&2z&y\\
288: z&2y&3w&2z\\
289: 0&z&0&3w
290: \end{pmatrix}
291: \]
292: we conclude that $D$ has equation $\det A=0$. 
293: After division by $3$ this determinant is
294: \[
295: -y^2z^2+4wy^3+4xz^3-18wxyz+27w^2x^2.
296: \]
297: In fact each of the columns of this matrix determines a vector field in $\Der(-\log D)$; 
298: for the group $\Gl_2(\CC)$ acts linearly on $B_{2,3}$ by composition on the right, and, up to a sign, the four columns here are the infinitesimal generators of this action corresponding to a basis of $\gl_2(\CC)$. 
299: Each is tangent to $D$, since the action preserves $D$. 
300: 
301: Further examples of irreducible linear free divisors can be found (though not under this name) in the paper \cite{SK77} of Sato and Kimura. 
302: Besides our example, two, of ambient dimension $12$ and $40$, are described in \cite[\S5, Prop.~11, 15]{SK77}, and by repeated application of castling transformations, cf.~\cite[\S2]{SK77}, it is possible to generate infinitely many more, of higher dimensions. 
303: \end{asparaenum}
304: \end{exems}
305: 
306: In Section~\ref{38} of this paper we describe a number of further examples of 
307: linear free divisors, and in Section~\ref{48} we prove some results about linear bases for the module $\Ga(\CC^n,\Der(-\log D))$, and go on to classify all linear free divisors in dimension $n\le 4$.
308: 
309: Linear free divisors provide a new insight into a conjecture of H.~Terao \cite[Conj.~3.1]{Ter78} relating the cohomology of the complement of certain divisors $D$ 
310: to the cohomology of the complex $\Om^\bu(\log D)$ of forms with logarithmic poles along $D$.
311: For linear free divisors, the link between the complex $\Ga(\CC^n,\Om^\bu(\log D))$ and $H^*(\CC^n\ssm D)$ can be understood as follows.
312: 
313: \begin{defi}\label{17}
314: For a linear free divisor $D$ defined by $\Delta\in\CC[x]$, we consider the subgroup
315: \[
316: G_D:=\{A\in\Gl_n(\CC)\mid A(D)=D\}=\{A\in\Gl_n(\CC)\mid\Delta\circ A\in\CC\cdot\Delta\}
317: \]
318: with identity component $G_D^\circ$ and Lie algebra $\gg_D$.
319: We call $D$ \emph{reductive} if $G_D^\circ$, or equivalently $\gg_D$, is reductive.
320: \end{defi}
321: 
322: It turns out that $\CC^n\ssm D$ is a single orbit of $G^\circ_D$ with finite isotropy group, so $H^*(\CC^n\ssm D;\CC)$ is isomorphic to the cohomology of $G^\circ_D$; this is explained in Section~\ref{8}. 
323: Moreover, $H^*(\Ga(\CC^n,\Om^\bu(\log D)))$ coincides with the Lie algebra cohomology of $\gg_D$ with complex coefficients.
324: For compact connected Lie groups $G$, a well-known argument shows that the Lie algebra cohomology coincides with the topological cohomology of the group. 
325: For linear free divisors the group $G^\circ_D$ is never compact, but the isomorphism also holds good for the larger class of reductive groups, and for a significant class of linear free divisors, $G^\circ_D$ is indeed reductive. 
326: In Section~\ref{15} we prove our main result:
327: 
328: \begin{theo}\label{3}
329: If $D$ is a reductive linear free divisor then
330: \beq\label{4}
331: H^*(\Ga(\CC^n,\Om^\bu(\log D)))\simeq H^*(\CC^n\smallsetminus D;\CC).
332: \eeq
333: \end{theo}
334: 
335: Among linear free divisors to which it applies are those arising as discriminants in representation spaces of quivers, as discussed in detail in \cite{BM06} and briefly in Section~\ref{31} below. 
336: 
337: Terao's conjecture remains open, though it has been answered in the affirmative for a very large class of arrangements in \cite{WY97}, using a technique developed in \cite{CMN96}. 
338: For general free divisors, a local result from which the global isomorphism of \eqref{4} follows holds when imposing the following additional hypothesis.
339: 
340: \begin{defi}\label{80}
341: A divisor $D$ is called \emph{quasihomogeneous} at $p\in D$ if the germ $(D,p)$ admits a local defining equation $\Delta\in\O_p$ that is weighted homogeneous with respect to weights $w_1,\dots,w_n\in\QQ_+$ in some local coordinate system $x_1,\dots,x_n$ centred at $p$.
342: Dividing $w_1,\dots,w_n$ by the weighted degree of $\Delta$, note that the preceding condition means that $\chi(\Delta)=\Delta$ where $\chi=\sum_{i=1}^nw_ix_i\p_i\in\Der(-\log D)_p$.
343: %By quasihomogeneity we mean quasihomogeneity at $0\in\CC^n$.
344: $D$ is called \emph{locally quasihomogeneous} if it is quasihomogeneous at $p$ for all $p\in D$.
345: We say \emph{homogeneous} instead of \emph{quasihomogeneous} if $w=1,\dots,1$.
346: \end{defi}
347: 
348: \begin{theo}[\cite{CMN96}]\label{6}
349: Let $D\subseteq\CC^n$ be a locally quasihomogeneous free divisor, let $U=\CC^n\smallsetminus D$, and let $j:U\to\CC^n$ be inclusion.
350: Then the de~Rham morphism
351: \beq\label{7}
352: \Om_X^{\bullet}(\log D)\to\mathbf{R}j_*\CC_U
353: \eeq
354: is a quasi-isomorphism.
355: \end{theo}
356: 
357: Grothendieck's \emph{Comparison Theorem} \cite{Gro66} asserts that a similar quasi-isomorphism holds for any divisor $D$, if instead of logarithmic poles we allow meromorphic poles of arbitrary order along $D$. 
358: Because of this similarity, we refer to the quasi-isomorphism of \eqref{7} as the Logarithmic Comparison Theorem (LCT) and to the global isomorphism \eqref{4} as the Global Logarithmic Comparison Theorem (GLCT).
359: Several authors have further investigated the range of validity of LCT, and established interesting links with the theory of $\D$-modules, in particular in
360: \cite{CN02}, \cite{CU05}, \cite{GS06}, \cite{Tor04}, and \cite{Wal05}.
361: 
362: Local quasihomogeneity was introduced in \cite{CMN96} as a technical device to make possible an inductive proof of the isomorphism in~\ref{6}.
363: Subsequently it turned out to have a deeper connection with the theorem. 
364: In particular by \cite{CCMN02}, for plane curves the logarithmic comparison theorem holds if and only if all singularities are quasihomogeneous.
365: The situation in higher dimensions remains unclear. 
366: There is as yet no counterexample to the conjecture that LCT is equivalent to the following weaker condition.
367: 
368: \begin{defi}\label{81}
369: A divisor $D$ is called \emph{Euler homogeneous} at $p\in D$ if there is a germ of vector field $\chi\in\mm_p\cdot\Der_p$ such that $\chi(\Delta)=\Delta$ for some local defining equation $\Delta\in\O_p$ of the germ $(D,p)$.
370: In this case, $\chi$ is called an \emph{Euler vector field} for $D$ at $p$.
371: %By \emph{Euler homogeneity} we mean Euler homogeneity at $p\in\CC^n$.
372: $D$ is called \emph{strongly Euler homogeneous} if it is Euler homogeneous at $p$ for all $p\in D$.
373: \end{defi}
374: 
375: \begin{rema}
376: The Euler homogeneity of $D$ is independent of the choice of an equation. 
377: If $\chi$ is an Euler vector field at $p$ for $D$ defined by $\Delta\in\O_p$, and $u\in\O_p^*$ is a unit, then the defining equation $u\Delta$ of $D$ at $p$ satisfies an equation
378: \[
379: (\chi(u)+u)^{-1}u\chi(u\Delta)=(\chi(u)+u)^{-1}(\chi(u)+u)u\Delta=u\Delta
380: \]
381: with Euler vector field $(\chi(u)+u)^{-1}u\chi$.
382: \end{rema}
383: 
384: In Section \ref{5} we examine the examples described in Sections \ref{38} and \ref{48} with respect to local quasihomogeneity and strong Euler homogeneity. 
385: It turns out that all linear free divisors in dimension $n\le4$ are locally quasihomogeneous and there is no linear free divisor which we know not to be strongly Euler homogeneous.
386: The optimistic reader could therefore conjecture that all linear free divisors are strongly Euler homogeneous, and also fulfil LCT and so also GLCT.
387: We do not know any counter-example to these statements.
388: 
389: In Subsection~\ref{62} we give examples of quivers $Q$ and dimension vectors $\d$ for which the discriminant in $\Rep(Q,\d)$ is a linear free divisor but is not locally quasihomogeneous. 
390: In such cases Theorem~\ref{6} therefore does not apply, but Theorem~\ref{3} does.
391: 
392: In Subsection~\ref{65}, we show that a linear free divisor does not need to be reductive for LCT to hold.
393: However we do not know whether reductiveness of the group implies LCT for linear free divisors.
394: The property of being a linear free divisor is not local, and our proof of GLCT here is quite different from the proof of LCT in \cite{CMN96}. 
395: 
396: The fact that linear free divisors in $\CC^n$ arise as the complement of the open orbit of an $n$-dimensional connected algebraic subgroup of $\Gl_n(\CC)$, means that there is some overlap between the topic of this paper and of the paper \cite{SK77}, where Sato and Kimura classify irreducible \emph{prehomogeneous vector spaces}, that is, triples $(G,\rho,V)$, where $\rho$ is an irreducible representation of the algebraic group $G$ on $V$, in which there is an open orbit. 
397: However, the hypothesis of irreducibility means that the overlap is slight. 
398: Any linear free divisor arising as the complement of the open orbit in an irreducible prehomogeneous vector space is necessarily irreducible by \cite[\S4, Prop.~12]{SK77}, whereas among our examples and in our low-dimensional classification (in Section~\ref{48}) all the linear free divisors except one (Example~\ref{1}(\ref{1b})) are reducible.
399: Even where $G$ is reductive, the passage from irreducible to reducible representations in this context is by no means trivial, including as it does substantial parts of the theory of representations of quivers.
400:   
401: 
402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
403: \section{Linear free divisors and subgroups of $\Gl_n(\CC)$}\label{8}
404: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
405: 
406: A degree zero vector field $\delta\in\Der$ can be identified with an $n\times n$ matrix $A=(a_{i,j})_{i,j}\in\CC^{n\times n}$ by $\delta=\sum_{i,j}x_ia_{i,j}\p_j=xA\p^t$.
407: Under this identification, the commutator of square matrices corresponds to the Lie bracket of vector fields.
408: 
409: Let $D\subseteq\CC^n$ be a reduced divisor defined by a homogeneous polynomial $\Delta\in\CC[x]$ of degree $d$.
410: 
411: \begin{defi}\label{23}
412: We denote by 
413: \[
414: \L_D:=\{xA\p^t\mid xA\p^t(\Delta)\in\CC\cdot\Delta\}\subseteq\Ga(\CC^n,\Der(-\log D))
415: \]
416: the Lie algebra of degree zero global logarithmic vector fields.
417: \end{defi}
418: 
419: Recall from Definition~\ref{2}, that $D$ is linear free if $\L_D$ contains a basis of $\Der(-\log D)$, and recall $G_D^\circ$ from Definition~\ref{17}.
420: 
421: \begin{lemm}\label{9}
422: $G_D^\circ$ is an algebraic subgroup of $\Gl_n(\CC)$ and $\gg_D=\{A\mid xA^t\p^t\in\L_D\}$.
423: \end{lemm}
424: 
425: \begin{proof} 
426: Clearly $G_D$ is a subgroup of $\Gl_n(\CC)$ and defined by a system of polynomial (determinantal) equations.
427: Thus $G_D$ and hence also $G_D^\circ$ is an algebraic subgroup of $\Gl_n(\CC)$.
428: The Lie algebra of $G_D^\circ$ consists of all $n\times n$-matrices $A$ such that
429: \[
430: \Delta\circ(I+A\eps)=a(\eps)\cdot\Delta\in\CC[\eps]\cdot\Delta
431: \]
432: where $\CC[\eps]=\CC[t]/\ideal{t^2}\ni[t]=:\eps$.
433: Taylor expansion of this equation with respect to $\eps$ yields
434: \[
435: \Delta+\p(\Delta)\cdot A\cdot x^t\cdot\eps=(a(0)+a'(0)\cdot\eps)\cdot\Delta
436: \]
437: and hence $a(0)=1$ and, by transposing the $\eps$-coefficient, $xA^t\p^t\in\L_D$.
438: The argument can be reversed to prove the converse by setting
439: \[
440: a(\eps):=1+(xA^t\p^t(\Delta)/\Delta)\cdot\eps.
441: \]
442: \end{proof}
443: 
444: \begin{lemm}\label{10}
445: The complement $\CC^n\ssm D$ of a linear free divisor is an orbit of $G_D^\circ$ with finite isotropy groups.
446: \end{lemm}
447: 
448: \begin{proof}
449: For $p\in\CC^n$, the orbit $G_D^\circ\cdot p$ is a smooth locally closed subset of $\CC^n$ whose boundary is a union of strictly lower dimensional orbits, cf.~\cite[Prop.~8.3]{Hum75}.
450: The orbit map $G_D^\circ\to G_D^\circ\cdot p$ sends $I_n+A\eps$ to $p+pA^t\eps$ and induces a tangent map
451: \beq\label{11}
452: \gg_D\onto T_p(G_D^\circ\cdot p),\quad A\mapsto pA^t.
453: \eeq
454: For $p\not\in D$, $\Der(-\log D)(p)$ and hence also $\L_D(p)$ is $n$-dimensional.
455: Then by Lemma \ref{9} and \eqref{11} $T_pG_D^\circ\cdot p$ and hence $G_D^\circ\cdot p$ are $n$-dimensional which implies the finiteness of the isotropy group of $p$ in $G_D^\circ$.
456: As this holds for all $p\not\in D$, the boundary of $G_D^\circ\cdot p$ must be $D$ and then $G_D^\circ\cdot p=\CC^n\ssm D$.
457: \end{proof}
458: 
459: Reversing our point of view we might try to find algebraic subgroups $G\subseteq\Gl_n(\CC)$ that define linear free divisors. 
460: This requires by definition that $G$ is $n$-dimensional and connected and by Lemma \ref{10} that there is an open orbit.
461: The complement $D$ is then a candidate for a free divisor.
462: Indeed $D$ is a divisor: comparing with \eqref{11}, $D$ is defined by the \emph{discriminant determinant}
463: \[
464: \Delta=\det\begin{pmatrix}A_1x^t&\cdots&A_nx^t\end{pmatrix}
465: \]
466: where $A_1,\dots,A_n$ is a basis of the Lie algebra $\gg$ of $G$ and we denote by $f=\Delta_\text{red}$ the reduced equation of $D$.
467: As the entries of the defining polynomial are linear, $\Delta$ is a homogeneous polynomial of degree $n$.
468: Thus, if $\Delta$ is not reduced, $D$ can not be linear free.
469: We shall see examples where this happens in the next section.
470: On the other hand, Saito's criterion \cite[Lem.~1.9]{Sai80} shows the following.
471: 
472: \begin{lemm}\label{12} 
473: Let the $n$-dimensional algebraic group $G$ act linearly on $\CC^n$ with an open orbit.
474: If $\Delta$ is reduced then $D$ is a linear free divisor.\qed
475: \end{lemm}
476: 
477: As a first step towards our main result, we now describe the cohomology of $\CC^n\ssm D$ in terms of $G_D^\circ$.
478: 
479: \begin{prop}\label{13}
480: Suppose that $D\subseteq\CC^n$ is a linear free divisor and let $G^\circ_{D,p}$ be the (finite) isotropy group of $p\in\CC^n\ssm D$ in $G_D^\circ$.
481: Then
482: \[
483: H^*(\CC^n\ssm D;\CC)=H^*(G^\circ_D;\CC)^{G^\circ_{D,p}}=H^*(G^\circ_D;\CC).
484: \] 
485: \end{prop}
486: 
487: \begin{proof}
488: By Lemma \ref{10}, $\CC^n\ssm D\cong G^\circ_D/G^\circ_{D,p}$ with finite $G^\circ_{D,p}$ and the first equality follows.
489: The second equality holds because $G^\circ_D$ is path connected, which means that left translation by $g\in G^\circ_{D,p}$ is homotopic to the identity and thus induces the identity map on cohomology.
490: \end{proof}
491: 
492: \begin{rema}\label{14} 
493: The argument for the second equality also shows that if $G^\circ_D$ is a finite quotient of the connected Lie group $G$ then $H^*(\CC^n\ssm D;\CC)\simeq H^*(G;\CC).$
494: We will use this below in calculating the cohomology of $\CC^n\ssm D$.
495: \end{rema}
496: 
497: 
498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499: \section{Cohomology of the complement and Lie algebra cohomology}\label{15}
500: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501: 
502: Let $\gg$ be a Lie algebra. 
503: The complex of Lie algebra cochains with coefficients in the complex representation $V$ of $\gg$ has $k$th term $\bigwedge^k_{\CC}\Hom_{\CC}(\gg,V)\cong \Hom_{\CC}(\bigwedge^k_{\CC}\gg,V)$, and differential $d_L:\bigwedge^k_\CC\Hom(\gg,V)\to\bigwedge^{k+1}_\CC\Hom(\gg,V)$ defined by
504: \begin{gather}\label{16}
505: (d_L\om)(v_1\wedge\dots\wedge v_{k+1})=\\
506: \nonumber\sum_{i<j}(-1)^{i+j}\om([v_i,v_j]\wedge v_1\dots\wedge\widehat{v_i}\wedge\dots\wedge\widehat{v_j}\wedge\dots\wedge v_{k+1})+\\
507: \nonumber\sum_i(-1)^{i+1}v_i\cdot\om(v_1\wedge\dots\wedge\widehat{v_i}\wedge\dots\wedge v_{k+1}).
508: \end{gather}
509: The cohomology of this complex is the \emph{Lie algebra cohomology} of $\gg$ with coefficients in $V$ and will be denoted $H^*_A(\gg;V)$.
510: 
511: The exterior derivative of a differential $k$-form satisfies an 
512: identical formula:
513: \begin{gather}\label{71}
514: d\om(\chi_1\wedge\dots\wedge\chi_{k+1})=\\
515: \nonumber\sum_{i<j}(-1)^{i+j}\om([\chi_i,\chi_j]\wedge\chi_1\wedge\dots\wedge\widehat{\chi_i}\wedge\dots\wedge\widehat{\chi_j}\wedge\dots\wedge\chi_{k+1})+\\
516: \nonumber\sum_i(-1)^{i+1}\chi_i\cdot\om(\chi_1\wedge\dots\wedge\widehat{\chi_i}\wedge\dots\wedge\chi_{k+1}).
517: \end{gather}
518: Here the $\chi_i$ are vector fields. 
519: 
520: When $D$ is a free divisor and $V=\O_p$ for some $p\in D$, it is tempting to conclude from the comparison of \eqref{16} and \eqref{71} that the 
521: complex $\Om^\bu(\log D)$ coincides with the complex of Lie algebra cohomology, with coefficients in $\O_p$, of the Lie algebra $\Der(-\log D)_p$. 
522: For $\Om^1(\log D)_p$ is the dual of $\Der(-\log D)_p$, and $\Om^k(\log D)=\bigwedge^k\Om^1(\log D)$.
523: However, this identification is incorrect, since, in the complex $\Om^\bu(\log D)$, both exterior powers and $\Hom$ are taken over the ring of coefficients $\O$, rather than over $\CC$, as in the complex of Lie algebra cochains. 
524: The cohomology of $\Om^\bu(\log D)_p$ is instead the \emph{Lie algebroid} cohomology of $\Der(-\log D)_p$ with coefficients in $\O_p$.
525: Nevertheless, when $D$ is a linear free divisor, there is the following important
526: link between these two complexes.
527: 
528: Recall $\L_D$ from Definition~\ref{23}. 
529: \begin{lemm}\label{82}
530: Let $D$ be a linear free divisor.
531: The complex $\Ga(\CC^n,\Om^\bu(\log D))_0$ of global homogeneous differential forms of degree zero coincides with the complex $\bigwedge^\bu_\CC\Hom(\L_D,\CC)$ of Lie algebra cochains with coefficients in $\CC$. 
532: \end{lemm}
533: 
534: \begin{proof}
535: First we establish a natural isomorphism between the corresponding
536: terms of the two complexes. 
537: We have
538: \begin{align*}
539: \Om^1(\log D)&=\Hom_{\O}(\Der(-\log D),\O)\\
540: &=\Hom_{\O}(\L_D\otimes_{\CC}\O,\O)\\
541: &=\Hom_{\CC}(\L_D,\CC)\otimes_{\CC}\O.
542: \end{align*}
543: Since $\Hom_{\CC}(\L_D,\CC)$ is purely of degree zero, and the degree zero
544: part of $\O$ consists just of $\CC$, the degree zero part of $\Ga(\CC^n,\Om^1(\log D))$ is
545: \[
546: \Ga(\CC^n,\Om^1(\log D))_0=\Hom_{\CC}(\L_D,\CC).
547: \]
548: Since moreover $\Ga(\CC^n,\Om^1(\log D))$ has no part of negative degree, it follows that 
549: \[
550: \Ga(\CC^n,\Om^k(\log D))_0
551: =\Ga\bigl(\CC^n,\bigwedge^k_{\O}\Om^1(\log D)\bigr)_0 
552: =\bigwedge^k_{\CC}\Hom_{\CC}(\L_D,\CC).
553: \]
554: Next, we show that the coboundary operators are the same. 
555: Because we are working with constant coefficients, the second sum on the right in \eqref{16} vanishes. 
556: Let $\chi_1,\dots,\chi_{k+1}\in\L_D$.
557: Then for $\om\in\Ga(\CC^n,\Om^k(\log D))_0$ and $i\in\{1,\dots,k+1\}$, $\om(\chi_1\w\dots\w\widehat{\chi_i}\w\dots\w \chi_{k+1})$ is a constant. 
558: It follows that the second sum on the right in \eqref{71} vanishes. 
559: Thus, the coboundary operator $d_L$ and the exterior derivative $d$ coincide.
560: \end{proof}
561: 
562: More generally let us consider weights $w=w_1,\dots,w_n\in\QQ_+$ and assign the weight $w_i$ (resp.\ $-w_i$) to $x_i$ and $dx_i$ (resp.\ to $\p_i$). 
563: Then the set of homogeneous vector fields or differential forms of a given degree is well defined. 
564: 
565: \begin{lemm}\label{83}
566: Suppose that the divisor $D\subseteq\CC^n$ is quasihomogeneous with respect to weights $w=w_1,\dots,w_n\in\QQ_+$.   
567: Then the following holds for any open set $U\subseteq\CC^n$:
568: \begin{enumerate}
569: \item\label{83a} If $\om\in\Ga(U,\Om^k(\log D))$ is $w$-homogeneous, then $L_\chi(\om)=\deg_w(\om)\om$, where $L_\chi$ is the Lie derivative with respect the Euler vector field $\chi=\sum_{i=1}^nw_ix_i\p_i$.
570: \item\label{83b} For any closed $\om\in \Ga(U,\Om^k(\log D))$ with decomposition $\om=\sum_{j\geq j_0}\om_j$ into $w$-homogeneous parts, $\om-\om_0$ is exact.
571: \item\label{83c} If $\Ga(U,\Om^k(\log D))_r\subseteq\Ga(U,\Om^k(\log D))$ denotes the subspace of $w$-homogeneous forms of $w$-degree $r$, then
572: \[
573: \Ga(U,\Om^\bu(\log D))_0\into \Ga(U,\Om^\bu(\log D))
574: \]
575: is a quasi-isomorphism.
576: \end{enumerate}
577: \end{lemm}
578: 
579: \begin{proof}\
580: \begin{asparaenum}\pushQED{\qed}
581: \item is a straightforward calculation, using Cartan's formula
582: $L_{\chi}(\om)=d\iota_{\chi}\om+\iota_{\chi}d\om$, where $\iota_{\chi}$ is
583: contraction by $\chi$. 
584: \item follows, for, if $\om$ is closed, so is $\om_j$ for
585: every $j$, and thus
586: \[
587: \om-\om_0=
588: \sum_{0\neq j\geq j_0}\om_j=
589: L_\chi\bigl(\sum_{0\neq j\geq j_0}\frac{\om_j}{j}\bigr)=
590: d(\iota_\chi\bigl(\sum_{0\neq j\geq j_0}\frac{\om_j}{j}\bigr)).
591: \]
592: \item is now an immediate consequence.\qedhere
593: \end{asparaenum}
594: \end{proof}
595: 
596: From Lemma~\ref{82} and Lemma~\ref{83}\eqref{83c} applied to $U=\CC^n$ we deduce the following
597: 
598: \begin{prop}\label{84}
599: Let $D\subseteq\CC^n$ be a linear free divisor. 
600: Then 
601: \[\pushQED{\qed}
602: H^*(\Ga(\CC^n, \Om^\bu(\log D)))\cong H^*_A(\L_D;\CC).\qedhere
603: \]
604: \end{prop}
605: 
606: Recall $G_D^\circ$ and $\gg_D$ from Definition~\ref{17}.
607: From Propositions~\ref{13} and \ref{84} we deduce
608: 
609: \begin{coro}\label{18}
610: The global logarithmic comparison theorem holds for a linear free divisor $D$ if and only if
611: \beq\label{19}\pushQED{\qed}
612: H^*(G_D^\circ;\CC)\cong H^*_{A}(\gg_D;\CC).\qedhere
613: \eeq
614: \end{coro}
615: 
616: There is such an isomorphism if $G$ is a connected compact real Lie group with Lie algebra $\gg$ (which is not our situation here). 
617: Left translation around the group gives rise to an isomorphism of complexes
618: \[
619: T:\bigwedge^\bullet\gg^*\to\bigl(\Om^\bu(G)^G,d\bigr)
620: \]
621: where $\gg^*=\Hom_{\RR}(\gg,\RR)$ and $\Om^\bu(G)^G$ is the complex of left-invariant real-valued differential forms on $G$. 
622: Composing this with the inclusion 
623: \beq\label{20} 
624: \bigl(\Om^\bu(G)^G,d\bigr)\to \bigl(\Om^\bu(G),d\bigr)\eeq
625: and taking cohomology gives a morphism 
626: \beq\label{21}
627: \tau_G\colon H^*_A(\gg_D;\RR)\to H^*(G;\RR).
628: \eeq
629: If $G$ is compact, \eqref{21} is an isomorphism. 
630: For from each closed $k$-form $\om$ we obtain a left-invariant closed $k$-form $\om_A$ by averaging:
631: \[
632: \om_A:=\frac{1}{|G|}\int_G\ell_g^*(\om)d\mu_L,
633: \]
634: where $\mu_L$ is a left-invariant measure and $|G|$ is the volume of $G$ with respect to this measure.  
635: As $G$ is path-connected, for each $g\in G$ $\ell_g$ is homotopic to the identity, so $\om$ and $\ell_g^*(\om)$ are equal in cohomology. 
636: It follows from this that $\om$ and $\om_A$ are also equal in cohomology.
637: 
638: Of course, this does not apply directly in any of the cases discussed here, since $G_D$ is not compact. 
639: Nevertheless if $G_D$ is a reductive group, the complexified morphism \eqref{21} is an isomorphism.
640: We now briefly outline the necessary definitions. 
641: Let $G_0$ be a compact Lie group. 
642: Then (\cite[\S 5.4, Thm.~10]{OV90}) $G_0$ has a faithful real representation. 
643: It follows (\cite[\S 3.4, Thm.~5]{OV90}) that $G_0$ has an affine real algebraic group structure. 
644: This allows its complexification. 
645: 
646: \begin{defi}\label{25}\
647: \begin{enumerate} 
648: \item The complex Lie algebra representation is {\em reductive} if it is the direct sum of a semi-simple ideal and a diagonalizable ideal.
649: \item The complex linear algebraic group $G$ is {\em reductive} if its Lie algebra (representation) is reductive.
650: \end{enumerate}
651: \end{defi}
652: 
653: The term ``reductive'' is due to the fact that these groups are characterised, among complex algebraic groups, by the complete reducibility of every finite-dimensional complex representation.  
654: Chapter 5 of \cite{OV90} establishes a bijection between compact Lie groups and reductive complex linear algebraic groups: 
655: 
656: \begin{theo}[{\cite[\S5.2, Thm.~5]{OV90}}]
657: On any compact Lie group $K$ there exists a unique real algebraic group structure, whose complexification $K(\CC)$ is reductive.
658: Any reductive complex algebraic group possesses an algebraic compact real form (of which it is therefore the complexification).\qed
659: \end{theo} 
660: 
661: The significance of this notion for us derives from the following fact:
662: 
663: \begin{theo}[{\cite[\S5.2 Thm.~2]{OV90}}]\label{89}
664: Let $G$ be a complex reductive algebraic group with an $n$-dimensional compact real form $K$. 
665: Then $G$ is diffeomorphic to $K\times\RR^n$.\qed
666: \end{theo}
667: 
668: \begin{coro}\label{24} 
669: If $G$ is a connected reductive complex algebraic group with complex Lie algebra $\gg$ then 
670: \[
671: H^*_A(\gg;\CC)\simeq H^*(G;\CC).
672: \]
673: \end{coro}
674: 
675: \begin{proof}
676: Let $K$ be a compact real form of $G$.
677: By \ref{89}, inclusion of $K$ into $G=K(\CC)$ induces an ismorphism on cohomology. 
678: The Lie algebra $\gg$ of $G$ is the complexification of the Lie algebra $\mathfrak{k}$ of $K$, so we have 
679: \[
680: H^*_A(\gg;\CC)\simeq H^*(\mathfrak{k};\RR)\otimes\CC\simeq H^*(K;\RR)\otimes\CC\simeq H^*(K;\CC)\simeq H^*(G;\CC),
681: \]
682: where the second isomorphism comes from the isomorphism (\ref{21}).
683: \end{proof}
684: 
685: From Corollary~\ref{18}, Definition~\ref{25}, and Corollary~\ref{24} we now conclude Theorem ~\ref{3} as announced in the introduction: the Global Logarithmic Comparison Theorem holds for all reductive linear free divisors.
686: 
687: Using the reductiveness of the group $\Gl_n(\CC)$, we will show in the next section that the group $G_D$ is reductive for divisors obtained as discriminants in the representation spaces of quivers.
688: The subgroup $B_n\subseteq\Gl_n(\CC)$ of upper triangular matrices is not reductive, and appears as the group $G_D$ in Example~\ref{40} which shows that reductivity is not necessary for the GLCT to hold. 
689: 
690: 
691: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
692: \section{Linear free divisors in quiver representation spaces}\label{31}
693: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
694: 
695: The following discussion summarises part of \cite{BM06}. 
696: A quiver $Q$ is a finite connected oriented graph; it consists of a set $Q_0$ of nodes and a set $Q_1$ of arrows joining some of them. 
697: For each arrow $\vp\in Q_1$ we denote by $t\vp$ (for ``tail'') and $h\vp$ (for ``head'') the nodes where it starts and finishes.
698: A (complex) representation $V$ of $Q$ is a choice of complex vector space $V_\alpha$ for each node $\alpha\in Q_0$ and linear map $V(\vp):V_{t\vp }\to V_{h\vp}$ for each arrow $\vp\in Q_1$.
699: For a fixed \emph{dimension vector}
700: \[
701: \d=(d_\alpha)_{\alpha\in Q_0}:=(\dim V_\alpha)_{\alpha\in Q_0}.
702: \]
703: and a choice of bases for the $V_\alpha$, $\alpha\in Q_0$, the \emph{representation space} of the quiver $Q$ of dimension $\d$ is
704: \[
705: \Rep(Q,\d):=\prod_{\vp\in Q_1}\Hom(\CC^{d_{t\vp}},\CC^{d_{h\vp}})\cong\prod_{\vp\in Q_1}\Hom(V_{t\vp},V_{h\vp}).
706: \] 
707: On this space the \emph{quiver group}
708: \[
709: \Gl(Q,\d):=\prod_{\alpha\in Q_0}\Gl_{d_\alpha}(\CC)\cong\prod_{\alpha\in Q_0}\Gl(V_{\alpha})
710: \]
711: acts, by
712: \beq\label{32}
713: \bigl((g_{\alpha})_{\alpha\in Q_0}\cdot V\bigr)_{\vp}:=g_{h\vp}V(\vp)g_{t\vp}^{-1}.
714: \eeq
715: This action factors through the group 
716: \beq\label{33}
717: Z:=\CC^*\cdot(I_{d_\alpha})_{\alpha\in Q_0}\subseteq\Z(\Gl(Q,\d))
718: \eeq
719: in the center of $\Gl(Q,\d)$ where $I_{d_\alpha}\in\Gl_{d_\alpha}(\CC)$ is the unit matrix.
720: The group $\Gl(Q,\d)/Z$ is reductive as, choosing a vertex $x_0\in Q_0$, we can consider it as a central quotient 
721: \beq\label{34}
722: \Gl(Q,\d)/Z\cong
723: \Bigl(\Sl_{d_{x_0}}(\CC)\times\prod _{x\in Q_0\ssm\{x_0\}}\Gl_{d_x}(\CC)\Bigr)\Big/\bigl(\mu_{d_{x_0}}\cdot \prod I_{d_x}\bigr)
724: \eeq
725: where $\mu_k\subseteq\CC^*$ denotes the cyclic subgroup of order $k$.
726: It acts faithfully on $\Rep(Q,\d)$.
727: For $\Rep(Q,\d)$ and $\Gl(Q,\d)/Z$ to play the role of $\CC^n$ and $G_D$ as in Section~\ref{8}, we must require
728: \begin{align}\label{35}
729: \sum_{n\in N}d_n^2-\sum_{\vp\in A}d_{t\vp}d_{h\vp}
730: &=\dim_{\CC}\Gl(Q,\d)-\dim_{\CC}\Rep(Q,\d)\\
731: \nonumber&=\dim Z
732: =1.
733: \end{align}
734: But this equality is not yet sufficient: 
735: it is also necessary that $\Gl(Q,\d)/Z$ has an open orbit.
736: This occurs if the general representation in $\Rep(Q,\d)$ is indecomposable. 
737: If both this last condition and \eqref{35} hold, $\d$ is called a \emph{real Schur root} of $Q$. 
738: In this case, there is a single open orbit, and the \emph{discriminant determinant} $\Delta$ defines its complement $D$, a divisor called the \emph{discriminant}. 
739: This is the consequence of a result due to Kraft and Riedtmann \cite[\S 2.6]{KR86}, which asserts that if the general representation is indecomposable it has only scalar endomorphisms.
740: Then
741: \beq\label{36}
742: \Gl(Q,\d)/Z\cong G_D=G_D^\circ.
743: \eeq
744: The above discussion combined with Theorem~\ref{3} proves the following 
745: 
746: \begin{theo}\label{37}
747: If $\d$ is a real Schur root of a quiver $Q$ and the discriminant $D$ in $\Rep(Q,\d)$ is reduced then $D$ is a linear free divisor that satisfies the GLCT.
748: \end{theo}
749: 
750: In \cite{BM06} it is shown that if, moreover, $Q$ is a \emph{Dynkin quiver}, i.e.\ its underlying unoriented graph is a Dynkin diagram of type $A_n$, $D_n$, $E_6$, $E_7$ or $E_8$, then $\Delta$ is always reduced, and thus defines a linear free divisor.
751: The significance of the Dynkin quivers is, that by a theorem of Gabriel \cite{Gab72}, they are the quivers of \emph{finite type}, i.e.\ the number of $\Gl(Q,\d)$ orbits in $\Rep(Q,\d)$ is finite. 
752: It is this that guarantees that $\Delta$ is always reduced, cf.~\cite[Prop.~5.4]{BM06}. 
753: It also implies that every root of a Dynkin quiver is a real Schur root.
754: 
755: \begin{coro}
756: If $\d$ is a (real Schur) root of a Dynkin quiver $Q$ then the discriminant $D$ in $\Rep(Q,\d)$ is a linear free divisor that satisfies GLCT.
757: \end{coro}
758: 
759: \begin{rema}
760: The argument showing that GLCT holds for the free divisors arising as discriminants in quiver representation spaces yields a simple topological proof of a theorem of V.~Kac \cite[p.~153]{Kac82} (see also \cite{Sch91}):
761: When $\d$ is a sincere (i.e.\ $\d_x>0$ for all $x\in Q_0$) real Schur root of a quiver $Q$ with no oriented cycles, the discriminant in $\Rep(Q,\d)$ has $|Q_0|-1$ irreducible components. 
762: The proof is this: 
763: the number of irreducible components of a divisor in a complex vector space is equal to the rank of $H^1$ of the complement.
764: From Theorem~\ref{3} we know that $H^1(\Rep(Q,\d)\ssm D;\CC)\simeq H^1_A(\gg_D;\CC)$; as by \eqref{34}
765: \[
766: \gg_D\simeq\sl_{d_{x_0}}(\CC)\oplus\bigoplus _{x\in Q_0\ssm\{x_0\}}\gl_{d_x}(\CC),
767: \]
768: it follows that
769: \[
770: H^1(\Rep(Q,\d)\ssm D;\CC)\simeq 0\oplus\bigoplus_{x\in Q_0\ssm\{x_0\}}H^1(\gl_{d_x}(\CC);\CC)
771: \]
772: and so has rank $|Q_0|-1$.
773: 
774: Another simple algebraic proof of Kac's theorem was pointed out to us by the referee.
775: It consists in determining the dimension of the vector space of rational function on $\CC^n$ with zeroes and poles along $D$ only and lifting them to the group $G_D$.
776: \end{rema}
777: 
778: 
779: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
780: \section{Examples of linear free divisors}\label{38}
781: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
782: 
783: The conclusion of Section~\ref{8} guides our search for linear free divisors. 
784: Our first example shows that the implication in Theorem~\ref{3} is not an equivalence.
785: 
786: We denote by
787: \beq\label{39}
788: E_{ij}=(\delta_{i,k}\cdot\delta_{j,l})_{k,l}\in\gl_n(\CC)
789: \eeq
790: the elementary matrix with $1$ in the $i$th row and $j$th column and $0$ elsewhere.
791: 
792: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
793: \subsection{A non-reductive example satisfying GLCT}
794: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
795: 
796: \begin{exem}\label{40}
797: For $n\ge2$, the group $B_n$ of $n\times n$ invertible upper triangular matrices is not reductive.
798: It acts on the space $\Sym_n(\CC)$ of symmetric matrices by transpose conjugation:
799: \[
800: B\cdot S=B^tSB.
801: \]
802: Under the corresponding infinitesimal action, the matrix $b$ in the Lie algebra $\bb_n$ gives rise to the vector field $\chi_b$ defined by
803: \[
804: \chi_b(S)=b^tS+Sb.
805: \]
806: The dimensions of $B_n$ and $\Sym_n(\CC)$ are equal. 
807: The discriminant determinant $\Delta$ is reduced and defines a linear free divisor $D=V(\Delta)$. 
808: \end{exem}
809: 
810: To see this, consider an elementary matrix $E_{ij}\in\bb_n$ and let $\chi_{ij}$ be the corresponding vector field on $\Sym_n(\CC)$. 
811: If $I$ is the $n\times n$ identity matrix, then $\chi_{ij}(I)=E_{ji}+E_{ij}$. 
812: The vectors $\chi_{ij}(I)$ for $1\leq i\leq j\leq n$ are therefore linearly independent, and $\Delta(I)\neq 0$.
813: 
814: For an $n\times n$ matrix $A$, let $A_j$ be the $j\times j$ matrix obtained by deleting the last $n-j$ rows and columns of $A$, and let $\det_j(A)=\det(A_j)$. 
815: If $B\in B_n$ and $S\in\Sym_n(\CC)$, then because $B$ is upper triangular, $(B^tSB)_j=B^t_jS_jB_j,$ and so $\det_j(B^tSB)=\det_j(B_j)^2\det_j(S)$. 
816: It follows that the hypersurface $D_j:=\{\det_j=0\}$ is invariant under the action, and the infinitesimal action of $B_n$ on $\Sym_n(\CC)$ is tangent to each. 
817: Thus $\Delta$ vanishes on each of them.
818: The sum of the degrees of the $D_j$ as $j$ ranges from $1$ to $n$ is equal to $\dim\Sym_n(\CC)$, and so coincides with the degree of $\Delta$. 
819: Hence $\Delta$ is reduced, and we conclude, by Lemma~\ref{12}, that 
820: $D=D_1\cup\cdots\cup D_n$ is a linear 
821: free divisor. 
822: In particular, when $n=2$, $D\subseteq\Sym_2(\CC)=\CC^3$ is the union of a quadric cone and one of its tangent planes.
823: 
824: We now give a proof that GLCT holds for $D$, in the spirit of the proofs of the preceding section, even though $D$ is not reductive. 
825: In fact LCT already follows, by Theorem~\ref{6}, from local quasihomogeneity, which we prove in Subsection~\ref{65} below.
826: 
827: \begin{prop}\label{73}
828: GLCT holds for the discriminant $D$ of the action of $B_n$ on $\Sym_n(\CC)$ in Example \ref{40}.
829: \end{prop}
830: 
831: \begin{proof}
832: The group $G_D^\circ$ is a finite quotient of the group $B_n$ of upper-triangular matrices in $\Gl_n(\CC)$. 
833: There is a deformation retraction of $B_n$ to the maximal torus $T$ consisting of its diagonal matrices, and, with respect to the standard coordinates $a_{ij}$ on matrix space, it follows that $H^*(B_n)$ is isomorphic to the free exterior algebra on the forms $da_{ii}/a_{ii}$. 
834: Each of these is left-invariant, and it follows that the map $\tau_{B_n}\colon H^*_A(\bb_n;\CC)\to H^*(B_n;\CC)$ from \eqref{21} is an epimorphism. 
835: 
836: Similarly, the Lie algebra complex $\bigwedge^\bu\bb_n^*$ has a contracting homotopy to its semisimple part. 
837: We may consider it as the complex of left-invariant forms on the group $B_n$. 
838: Assign weights $w_1,\ldots,w_n$ to the columns and weights $-w_1,\ldots,-w_n$ to the rows. 
839: This gives the elementary matrix $E_{ij}\in\bb_n$ the weight $w_i-w_j$.
840: If $\eps_{i,j}\in\bb_n^*$ denotes the dual basis and we assign the weight $0$ to $\CC$ then $\wt(\eps_{i,j})=-\wt(E_{i,j})$.
841: With respect to the resulting gradings of $\bb_n$ and $\bb_n^*$, both the Lie bracket and the differential $d_L$ of the complex $\bigwedge^\bu\bb_n^*$ are homogeneous of degree 0, cf.~\eqref{16}.
842: 
843: Let $E=\sum_iw_iE_{ii}$, and let $\iota_E:\bigwedge^\bu\bb_n^*\to\bigwedge^\bu\bb_n^*$ be the operation of \emph{contraction by $E$} defined by
844: \[
845: (\iota_E\om)(v_1\wedge\cdots\wedge v_k):=\om(E\wedge v_1\wedge\cdots\wedge v_k).
846: \]
847: Observe that for each generator $E_{ij}\in\bb_n$ we have
848: \beq\label{78}
849: [E,E_{ij}]=(w_i-w_j)\cdot E_{ij}=\wt(E_{ij})\cdot E_{ij}.
850: \eeq
851: We claim that the operation
852: \[
853: L_E:=\iota_Ed_L+d_L\iota_E,
854: \]
855: of taking the Lie derivative along $E$ has the effect of multiplying each homogeneous element of $\bigwedge^\bu\bb_n^*$ by its $w$-degree. 
856: Indeed the operation $L_E$ is a derivation of degree zero on $\bigwedge^\bu\bb_n^*$, 
857: and the result on 1 forms,
858: \[
859: L_E(\eps_{i,j})=(w_j-w_i)\eps_{i,j},
860: \]
861: is therefore sufficient and can be easily checked by direct calculation.
862: 
863: Thus $L_E$ defines a contracting homotopy from $\bigwedge^\bu\mathfrak{b}_n^*$ to its $w$-degree $0$ part $\bigl(\bigwedge^\bu \mathfrak{b}_n^*\bigr)_0$, by exactly the same calculation as in Lemma \ref{83}, but with $\Ga(U,\Om^k(\log D))$ and $L_\chi$ replaced respectively by $\bigwedge^\bu\mathfrak{b}_n^*$ and $L_E$. 
864: If we choose $w_1<\cdots<w_n$ then all off-diagonal members of the basis $\{\eps_{i,j}\}_{1\le i\le j\leq n}$ of $\bb_n^*$ have strictly positive $w$-degree.
865: It follows that 
866: \[
867: \bigwedge^\bu\mathfrak{b}_n^*\simeq\Bigl(\bigwedge^\bu\mathfrak{b}_n^*\Bigr)_0=\bigwedge^\bu\ideal{\eps_{1,1},\dots,\eps_{n,n}}=\bigwedge^\bu\ttt^*
868: \]
869: where $\ttt$ is the Lie algebra of the torus $T$ above.
870: The differential $d_L$ is zero on this subcomplex, showing that $\tau_{B_n}$ is an isomorphism.
871: \end{proof}
872: 
873: 
874: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
875: \subsection{Discriminants of quiver representations}
876: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
877: 
878: The following example, due to Ragnar-Olaf Buchweitz, is of the type discussed in Section~\ref{31}.
879: 
880: \begin{exem}\label{41}
881: In the space $M_{n,n+1}(\CC)$ of $n\times(n+1)$ matrices, let $D$ be the divisor defined by the vanishing of the product of
882: the maximal minors. 
883: That is, for each matrix $A\in M_{n,n+1}(\CC)$, let $A_j$ be $A$ minus its $j$'th column, and let $\Delta_j(A)=\det(A_j)$. 
884: Then 
885: \[
886: D=\{A\in M_{n,n+1}(\CC):\delta=\prod_{i=1}^{n+1}{\Delta}_j(A)=0\}.
887: \]
888: It is a linear free divisor. 
889: Here, as the group $G$ in Remark~\ref{14} we may take the product $Gl_n(\CC)\times \{\diag(1,\lambda_1,\dots,\lambda_n):\lambda_1,\dots,\lambda_n\in\CC^*\}$, acting by
890: \[
891: \bigl(A,\diag(1,\lambda_1,\dots,\lambda_n)\bigr)\cdot M=
892: A\cdot M\cdot\diag(1,\lambda_1,\dots,\lambda_n)^{-1}
893: \]
894: The placing of the $1$ in the first entry of the diagonal matrices is
895: rather arbitrary; it could be placed instead in any other fixed position
896: on the diagonal.
897: That $D$ is a linear free divisor follows from the fact that
898: \begin{enumerate}
899: \item the complement of $D$ is a single orbit, so the discriminant determinant
900: is not identically zero, and
901: \item the degree of $D$ is equal to the dimension of $G_D$, so the discriminant
902: determinant is reduced.
903: \end{enumerate}
904: \end{exem}
905: 
906: In our Example~\ref{41}, $M_{n,n+1}(\CC)$ is the representation space of the star quiver consisting of one sink and $n+1$ sources, with dimension vector assigning dimension $n$ to the sink and $1$ to each of the sources.
907: The case $n=5$ is shown in Figure~\ref{29}.
908: \begin{figure}[h]
909: \caption{A star quiver with $1$ sink and $6$ sources and $\d=(5,1,\dots,1)$}\label{29}
910: \[
911: \xymat@C=.15in{&\bullet^1\ar[dr]&&\bullet^1\ar[dl]\\
912: \bullet^1\ar[rr]&&\bullet^5&&\bullet^1\ar[ll]\\
913: &\bullet^1\ar[ur]&&\bullet^1\ar[ul]}
914: \]
915: \end{figure}
916: Once we have chosen a basis for each $V_\alpha$, each $V(\vp)$ is represented by an $n\times 1$ matrix; together they make up an $n\times (n+1)$ matrix. 
917: So the basis identifies $\Rep(Q,\d)=M_{n,n+1}(\CC)$ and then
918: \beq\label{42}
919: \Gl(Q,\d)=\Gl_n(\CC)\times\Gl_1(\CC)^{n+1}
920: \eeq
921: and the action in \eqref{32} becomes
922: \beq\label{43}
923: (A,\lambda_1,\dots,\lambda_{n+1})\cdot M=AM\diag(\lambda_1^{-1},\dots,\lambda_{n+1}^{-1}).
924: \eeq
925: From \eqref{33}, \eqref{34}, and \eqref{36}, result isomorphisms
926: \beq\label{70}
927: \Gl(Q,\d)\cong G_D\times Z,\quad Z=\CC^*\cdot(I_n,(1,\dots,1)),
928: \eeq
929: defined by normalizing an arbitrary element in the second factor.
930: 
931: Many more examples of linear free divisors can be found by similar means in representation spaces of quivers. 
932: The next example, also from \cite{BM06}, is a non Dynkin quiver where finiteness of orbits fails and $\Delta$ is not reduced.
933: 
934: \begin{exem}\label{44}
935: Consider the star quiver of Example~\ref{41} with $n=3$ with $\d=(3,1,1,1,1)$, as before, and now reverse the direction of one of the arrows.
936: The result is shown in Figure~\ref{50}.
937: \begin{figure}[h]
938: \caption{A modified star quiver with $\d=(3,1,1,1,1)$}\label{50}
939: \[
940: \xymat{&\bullet^1\\
941: \bullet^1\ar[r]^B&\bullet^3\ar[u]_A&\bullet^1\ar[l]^D\\
942: &\bullet^1\ar[u]^C}
943: \]
944: \end{figure}
945: The four hypersurfaces $\det(AB)=0$, $\det(AC)=0$, $\det(AD)=0$, $\det(BCD)=0$, are invariant under the action of the subgroup $G_D\subseteq\Gl(Q,\d)$ of Example~\ref{41}. 
946: However, the last of these is made up of infinitely many orbits: 
947: if the images of $B$, $C$ and $D$ lie in a plane $P$, then together with $\ker(A\cap P)$ they determine a cross-ratio. 
948: The discriminant determinant is equal, up to a scalar factor, to 
949: \[
950: \Delta=\det(AB)\cdot\det(AC)\cdot\det(AD)\cdot(\det(BCD))^2.
951: \]
952: \end{exem}
953: 
954: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
955: \subsection{Incomplete collections of maximal minors}\label{45} 
956: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
957: 
958: In the space $M_{m,n}(\CC)$ of $m\times n$ matrices with $n>m+1$, the product of all of the maximal minors no longer defines a linear free divisor, by reason of its degree. 
959: However, certain collections of $n$ maximal minors do define free divisors. There is a simple procedure for generating infinitely many  such collections, first described in \cite{Nie05}:
960: 
961: The space $M_{m,n}(\CC)$ can still be viewed as $\Rep(Q,\d)$ where $Q$ is the star quiver of Example~\ref{41} with $1$ sink and $n$ sources, and $\d=(m,1,\dots,1)$. 
962: As before, the quiver group $\Gl(Q,\d)$ acts with $1$-dimensional kernel $Z$, but now
963: \[
964: \dim\Gl(Q,\d)-1=m^2+n-1<mn=\dim M_{m,n}(\CC),
965: \]
966: making an open orbit impossible.
967: Therefore we replace $\Gl(Q,\d)$ by a group
968: \beq\label{79}
969: G:=\Gl_m(\CC)\times G_R
970: \eeq
971: with $\dim G/Z=\dim M_{m,n}$ by augmenting the second factor in \eqref{42} to a group $G_R\subseteq\Gl_n(\CC)$ with $\dim  G_R=mn-m^2+1$. 
972: To construct $G_R$, we consider an auxiliary quiver $\tilde Q=(Q_0,Q_1)$ with $Q_0=\{1,\dots,n\}$ and $Q_1\subseteq Q_0^2$ satisfying the following conditions:
973: \begin{itemize}
974: \item $|Q_1|=mn-m^2+1$;
975: \item $(i,i)\in Q_1$ for all $i\in Q_0$;
976: \item $(i,j)\in Q_1$ and $(j,k)\in Q_1$ implies that $(i,k)\in Q_1$.
977: \end{itemize}
978: These conditions are exactly those we need for the following formula:
979: \beq\label{72}
980: G_R:=\CC^{Q_1}\ssm\{\det=0\}\subseteq\CC^{n\times n}\ssm\{\det=0\}=\Gl_n(\CC)
981: \eeq
982: to define a group.
983: We write $(Q_0,Q_1)=:Q(G_R)$.
984: This group $G_R$ is generated by $\Gl_1(\CC)^n=\diag(\CC^*,\dots,\CC^*)$ and $mn-m^2-n+1$ supplementary elementary matrices $I_n+\CC\cdot E_{i,j}$ with $(i,j)\in Q_1$ and $i\ne j$, cf.~\eqref{39}.
985: 
986: The action of $G$ on $M_{m,n}(\CC)$ extends that in \eqref{43} by right multiplication of $G_R$ and factors through $G/Z$ with $Z=\CC\cdot (I_m, I_n)$ which is, as in \eqref{34}, a central quotient
987: \beq\label{74}
988: G/Z\cong(\Sl_m(\CC)\times G_R)/\mu_m\cdot(I_m,I_n)
989: \eeq
990: where $\mu_k\subseteq\CC^*$ denotes the cyclic subgroup of order $k$. 
991: 
992: \begin{prop}\label{76}\
993: \begin{asparaenum}
994: \item If the discriminant determinant $\Delta$ of the action of $G$ is not identically zero and the action of $G$ preserves the divisors of zeros of precisely $n$ distinct $m\times m$ minors, then the union of these divisors is a linear free divisor $D=V(\Delta)$.
995: \item If the action of $G$ preserves the divisor of zeros of more than $n$ distinct $m\times m$ minors then $\Delta$ is identically zero.
996: \end{asparaenum}
997: \end{prop}
998: 
999: \begin{proof}
1000: Any algebraic set preserved by the action of $G$ is contained in $V(\Delta)$.
1001: By construction, if $\Delta$ is not identically zero then its degree is $mn$.
1002: If moreover the action of $G$ preserves the zero set of $n$ distinct $m\times m$ minors then $\Delta$ is reduced and Lemma~\ref{12} shows that $V(\Delta)$ is a linear free divisor. 
1003: \end{proof}
1004: 
1005: \begin{lemm}\label{46}
1006: Right multiplication by $I_n+\CC\cdot E_{i,j}$ preserves the divisor defined by an 
1007: $m\times m$ minor if and only if the minor either contains column $i$ of the generic matrix or does not contain column $j$.
1008: \end{lemm}
1009: 
1010: \begin{proof}
1011: Suppose that the $m\times m$ submatrix $M'$ of the generic $m\times n$ matrix $M$ contains column $j$ but not column $i$. 
1012: Let $p$ be a point of $\{\det M'=0\}$ at which $M'$ has rank $m-1$ and the matrix $M''$ obtained from $M'$ by replacing column $j$ by column $i$ has rank $m$. 
1013: Then $\det(M'\cdot(I_n+E_{i,j}))(p)\neq 0$. 
1014: That is, $\cdot(I_n+\CC\cdot E_{i,j})$ does not preserve $\{\det M'=0\}$. 
1015: Similarly, $\det M'$ is clearly invariant under $\cdot(I_n+\CC\cdot E_{i,j})$ if $M'$ contains both columns $i$ and $j$.
1016: \end{proof}
1017: 
1018: \begin{exem}\label{47}
1019: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1020: \begin{annalif}
1021: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1022: Let $m=3$ and $n=6$. 
1023: The quiver
1024: \[
1025: \xymatrix@C=3mm@R=2mm{
1026: 1\ar[dr]&&3\ar[dl]\ar[dr]&&5\ar[dl]\\
1027: &2&&4&}
1028: \]
1029: determines admissible minors $M_{123}$, $M_{345}$, $M_{135}$, $M_{136}$, $M_{156}$, $M_{356}$, and the associated divisor (the zero locus of their product) is a linear free divisor. 
1030: Other linear free divisors that can be constructed by these methods are listed, for small values of $m,n$, in the preprint version \cite{GMNS06} of this article.
1031: They are not in general reductive. 
1032: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1033: \end{annalif}
1034: \begin{arxiv}
1035: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1036: \begin{asparaenum}
1037: 
1038: \item Take $m=2$, $n=4$. 
1039: We refer to the $2\times 2$ submatrix of the generic matrix containing columns $i$ and $j$ as $M_{ij}$.
1040: We need one supplementary column operation type. 
1041: By reordering the columns we may assume this is $\cdot(I_n+\CC\cdot E_{3,4})$.
1042: The set of submatrices satisfying the condition of Lemma~\ref{46} is $M_{12},M_{13},M_{23},M_{34}$. 
1043: The product of their determinants therefore defines a linear free divisor. 
1044: The group $G_R$ consists of matrices of the form 
1045: \[
1046: \begin{pmatrix}
1047: 1&0&0&0\\
1048: 0&\lambda_2&0&0\\
1049: 0&0&\lambda_3&\lambda_{34}\\
1050: 0&0&0&\lambda_4
1051: \end{pmatrix}
1052: \]
1053: 
1054: \item\label{47b}Take $m=2$, $n=5$. 
1055: Then we require two supplementary generators which means that the quiver $Q(\tilde G_R)$ has two arrows.
1056: Our Table~\ref{85} shows all such quivers with no more than five vertices (up to reordering of columns).
1057: Note that the quiver
1058: \[
1059: \xymat{1\ar[r]&2\ar[r]&3}
1060: \]
1061: does not appear since the conditions on $Q_1$ force it to become 
1062: \[
1063: \xymat{1\ar[r]\ar@/^/[rr]&2\ar[r]&3}.
1064: \]
1065: The symbols $*$ and $\star$ denote arbitrary elements of $\CC^*$ and $\CC$ respectively. 
1066: 
1067: \begin{longtable}{|c|c|c|c|}
1068: \caption{LFDs from collections of minors for $m=2$, $n=5$}\label{85}\\
1069: \hline
1070: \rule{0pt}{11pt}$Q(\tilde G_R)$ & $\tilde G_R$ & Admissible minors & LFD?\\
1071: \hline
1072: $
1073: \xymat@R=1mm
1074: {&4\\
1075: 3\ar[ur]\ar[dr]&\\
1076: &5}
1077: $
1078: &
1079: $
1080: \begin{pmatrix}
1081: *&0&0&0&0\\
1082: 0&*&0&0&0\\
1083: 0&0&*&\star&\star\\
1084: 0&0&0&*&0\\
1085: 0&0&0&0&*
1086: \end{pmatrix}
1087: $
1088: &
1089: $
1090: M_{12},M_{13},M_{23},M_{34},M_{35}
1091: $
1092: &
1093: Yes
1094: \\
1095: \hline
1096: $\xymat@R=1mm
1097: {2\ar[r]&3\\
1098: 4\ar[r]&5}
1099: $
1100: &
1101: $
1102: \begin{pmatrix}
1103: *&0&0&0&0\\
1104: 0&*&\star&0&0\\
1105: 0&0&*&0&0\\
1106: 0&0&0&*&\star\\
1107: 0&0&0&0&*
1108: \end{pmatrix}
1109: $
1110: &
1111: $
1112: M_{12},M_{14},M_{23},M_{24},M_{45}
1113: $
1114: &
1115: Yes
1116: \\
1117: \hline
1118: $
1119: \xymat@R=1mm
1120: {3\ar[dr]&\\
1121: &5\\
1122: 4\ar[ur]&}
1123: $
1124: &
1125: $
1126: \begin{pmatrix}
1127: *&0&0&0&0\\
1128: 0&*&0&0&0\\
1129: 0&0&*&0&\star\\
1130: 0&0&0&*&\star\\
1131: 0&0&0&0&*
1132: \end{pmatrix}
1133: $
1134: &
1135: $
1136: M_{12},M_{13},M_{23},M_{14},M_{24},M_{34}
1137: $
1138: &
1139: No
1140: \\
1141: \hline
1142: $
1143: \xymat@R=1mm
1144: {1\ar@/^/[r]&\ar@/^/[l]2}
1145: $
1146: &
1147: $
1148: \begin{pmatrix}
1149: *&\star&0&0&0\\
1150: \star&*&0&0&0\\
1151: 0&0&*&0&0\\
1152: 0&0&0&*&0\\
1153: 0&0&0&0&*\\
1154: \end{pmatrix}
1155: $
1156: &
1157: $
1158: M_{12},M_{34},M_{35},M_{45}
1159: $
1160: &
1161: No
1162: \\
1163: \hline
1164: \end{longtable}
1165: 
1166: Only the first and second yield linear free divisors. 
1167: The fourth fails because although the group has the right dimension and an open orbit, the discriminant determinant is not reduced: 
1168: the minor $M_{12}$ divides it with multiplicity $2$.
1169: 
1170: \item Take $m=3$, $n=5$. Again we need two supplementary generators, and there are the same four cases as when $(m,n)=(2,5)$. 
1171: In each case the group $G_R$ is the same as in the previous example. 
1172: The Table~\ref{86} of those which give rise to free divisors is different from the previous example now.
1173: 
1174: \begin{longtable}{|c|l|l|}
1175: \caption{LFDs from collections of minors for $m=3$, $n=5$}\label{86}\\
1176: \hline
1177: \rule{0pt}{11pt}Admissible minors & LDF?\\
1178: \hline
1179: $M_{123},M_{134},M_{234},M_{345},M_{135},M_{235}$ & No\\
1180: \hline
1181: $M_{123},M_{124},M_{234},M_{145},M_{245}$ & Yes\\
1182: \hline
1183: $M_{123},M_{124},M_{134},M_{234},M_{345}$ & Yes\\
1184: \hline
1185: $M_{123},M_{124},M_{125},M_{345}$ & No\\
1186: \hline
1187: \end{longtable}
1188: 
1189: \item Take $m=3$, $n=6$. 
1190: We need four supplementary generators and show all possible quivers with no more than six vertices in Table~\ref{87}. 
1191: 
1192: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1194: \begin{longtable}{|c|c|l|c|}
1195: \caption{LFDs from collections of minors for $m=3$, $n=6$}\label{87}\\
1196: \hline
1197: \rule{0pt}{11pt}$Q(\tilde G_R)$ & Admissible minors & LFD?\\
1198: \hline
1199: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1200: \xymat@C=3mm@R=1mm 
1201: {2&&1\ar[ll]\ar[dl]\ar[dr]\ar[rr]&&5\\
1202: &3&&4&}
1203: &
1204: $
1205: \begin{matrix}
1206: M_{123},M_{124},M_{125},M_{134},M_{135},\\
1207: M_{145},M_{126},M_{136},M_{146},M_{156}                         
1208: \end{matrix}
1209: $
1210: &
1211: No
1212: \\
1213: \hline
1214: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1215: \xymat@C=3mm@R=1mm
1216: {&2\ar[dr]&&3\ar[dl]&\\
1217: 1\ar[rr]&&5&&4\ar[ll]}
1218: &
1219: $
1220: \begin{matrix}
1221: M_{123},M_{124},M_{134},M_{234},M_{126},\\
1222: M_{136},M_{146},M_{236},M_{246},M_{346}                         
1223: \end{matrix}
1224: $
1225: &
1226: No
1227: \\
1228: \hline
1229: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1230: \xymat@C=3mm@R=1mm
1231: {&1\ar[dl]\ar[dr]&&&4\ar[dl]\ar[dr]&\\
1232: 2&&3&5&&6}
1233: &
1234: $
1235: \begin{matrix}
1236: M_{123},M_{124},M_{134},M_{456},M_{145},M_{146}
1237: \end{matrix}
1238: $
1239: &
1240: Yes
1241: \\
1242: \hline
1243: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1244: \xymat@C=3mm@R=1mm
1245: {1\ar[dr]&&2\ar[dl]&4\ar[dr]&&5\ar[dl]\\
1246: &3&&&6}
1247: &
1248: $
1249: \begin{matrix}
1250: M_{123},M_{124},M_{125},M_{145},M_{245},M_{456}
1251: \end{matrix}
1252: $
1253: &
1254: Yes
1255: \\
1256: \hline
1257: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1258: \xymat@C=3mm@R=1mm
1259: {1\ar[dr]&&2\ar[dl]&4&&5\\
1260: &3&&&6\ar[ul]\ar[ur]}
1261: &
1262: $
1263: \begin{matrix}
1264: M_{123},M_{126},M_{146},M_{156},\\
1265: M_{246},M_{256},M_{456}
1266: \end{matrix}
1267: $
1268: &
1269: No
1270: \\
1271: \hline
1272: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1273: \xymat@C=6mm@R=2mm
1274: {1\ar[d]&3\ar[dr]&4\ar[d]&5\ar[dl]\\
1275: 2&&6}
1276: &
1277: $
1278: \begin{matrix}
1279: M_{123},M_{124},M_{125},M_{134},\\
1280: M_{135},M_{145},M_{345}
1281: \end{matrix}
1282: $
1283: &
1284: No
1285: \\
1286: \hline
1287: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1288: \xymat@C=6mm@R=2mm
1289: {1\ar[d]&&3\ar[dl]\ar[d]\ar[dr]\\
1290: 2&4&5&6}
1291: &
1292: $
1293: \begin{matrix}
1294: M_{123},M_{134},M_{135},M_{136},\\
1295: M_{345},M_{346},M_{356}
1296: \end{matrix}
1297: $
1298: &
1299: No
1300: \\
1301: \hline
1302: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1303: \xymat@C=3mm@R=2mm
1304: {1\ar[dr]&&3\ar[dl]\ar[dr]&&5\ar[dl]\\
1305: &2&&4&}
1306: &
1307: $
1308: \begin{matrix}
1309: M_{123},M_{345},M_{135},M_{136},M_{156},M_{356}
1310: \end{matrix}
1311: $
1312: &
1313: Yes
1314: \\
1315: \hline
1316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1317: \xymat@C=3mm@R=2mm
1318: {&1\ar[dl]\ar[dr]&&4\ar[dl]\ar[dr]&\\
1319: 2&&3&&5}
1320: &
1321: $
1322: \begin{matrix}
1323: M_{124},M_{126},M_{134},M_{456},M_{145},M_{146}
1324: \end{matrix}
1325: $
1326: &
1327: Yes
1328: \\
1329: \hline
1330: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1331: \xymat@R=3mm@C=6mm
1332: {1\ar[d]&3\ar[d]\ar[dr]&4\ar[d]\\
1333: 2&5&6}
1334: &
1335: $
1336: \begin{matrix}
1337: M_{123},M_{124},M_{134},M_{135},M_{345},M_{346}
1338: \end{matrix}
1339: $
1340: &
1341: Yes
1342: \\
1343: \hline
1344: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1345: \xymat@R=3mm@C=6mm
1346: {1\ar[d]&3\ar[dl]\ar[d]\ar[dr]&\\
1347: 2&4&5}
1348: &
1349: $
1350: \begin{matrix}
1351: M_{123},M_{134},M_{135},M_{136},\\
1352: M_{345},M_{346},M_{356}
1353: \end{matrix}
1354: $
1355: &
1356: No
1357: \\
1358: \hline
1359: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1360: \xymat@R=3mm@C=6mm
1361: {1\ar[d]\ar[dr]&2\ar[d]&3\ar[dl]\\
1362: 4&5}
1363: &
1364: $
1365: \begin{matrix}
1366: M_{123},M_{124},M_{134},M_{126},\\
1367: M_{136},M_{146},M_{236}
1368: \end{matrix}
1369: $
1370: &
1371: No
1372: \\
1373: \hline
1374: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1375: \xymat@R=3mm@C=6mm
1376: {1\ar[d]\ar[r]&4\\
1377: 3&2\ar[l]\ar[u]}
1378: &
1379: $
1380: \begin{matrix}
1381: M_{123},M_{124},M_{125},M_{126},M_{156},M_{256}
1382: \end{matrix}
1383: $
1384: &
1385: Yes
1386: \\
1387: \hline
1388: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1389: \xymat@R=3mm@C=6mm
1390: {1\ar[d]\ar[r]&2\ar[dl]&\\
1391: 3&4\ar[r]&5}
1392: &
1393: $
1394: \begin{matrix}
1395: M_{123},M_{124},M_{126},M_{145},M_{146},M_{456}
1396: \end{matrix}
1397: $
1398: &
1399: Yes
1400: \\
1401: \hline
1402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1403: \xymat@R=3mm@C=6mm
1404: {1\ar[d]\ar[r]&2\ar[dl]\\
1405: 3&4\ar[l]}
1406: &
1407: $
1408: \begin{matrix}
1409: M_{124},M_{125},M_{126},M_{145},\\
1410: M_{146},M_{156},M_{456}
1411: \end{matrix}
1412: $
1413: &
1414: No
1415: \\
1416: \hline
1417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1418: \xymat@R=3mm@C=6mm
1419: {&1\ar[dl]\ar[d]\ar[dr]&\\
1420: 2\ar[r]&3&4}
1421: &
1422: $
1423: \begin{matrix}
1424: M_{123},M_{124},M_{145},M_{146},\\
1425: M_{125},M_{126},M_{156}
1426: \end{matrix}
1427: $
1428: &
1429: No
1430: \\
1431: \hline
1432: \end{longtable}
1433: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1434: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1435: \end{asparaenum}
1436: 
1437: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1438: \end{arxiv}
1439: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1440: \end{exem}
1441: 
1442: \begin{prop}\label{75}
1443: Let $D=V(\Delta)$ be a linear free divisor as constructed above.
1444: If $Q(G_R)$ has no oriented loops then GLCT holds for $D$.
1445: \end{prop}
1446: 
1447: \begin{proof}
1448: As in the proof of Proposition~\ref{73}, one can show that
1449: \[
1450: \tau_{G_R}\colon H^*_A(\mathfrak{g}_R;\CC)\to H^*(G_R;\CC)
1451: \]
1452: from \eqref{21} is an isomorphism.
1453: Here the absence of oriented loops serves as a replacement for the upper triangularity in the preceding proof.
1454: Indeed, if there are no oriented loops in $Q( G_R)$, it is possible to order the
1455: vertices of $Q(G_R)$, and thus the rows of the matrices in $M_{m,n}$, so that $i<j$ whenever there is an arrow from $i$ to $j$. 
1456: This puts all of the matrices of $G_R$ into upper triangular form. 
1457: It follows both that $G_R$ has a deformation retraction to its maximal torus $T$ consisting of diagonal matrices, and that the same contracting homotopy as in the proof of \ref{73} shows that the inclusion 
1458: $\bigwedge^\bu\mathfrak{t}^*\to \bigwedge^\bu\mathfrak{g}_R^*$ is
1459: a homotopy equivalence. 
1460: Thus $H^*(T):H^*_A(\mathfrak{g}_R;\CC)\to H^*(G_R;\CC)$ is an isomorphism.
1461: 
1462: Also for $G=\Sl_m(\CC)$ the map $\tau_G$ from \eqref{21} is an isomorphism.
1463: So by applying the the K\"unneth formulas for both Lie algebra and complex cohomology, the same holds for $G=\Sl_m(\CC)\times G_R$. 
1464: By \eqref{74}, $G/Z$ is connected as a finite quotient of the connected group $\Sl_m(\CC)\times G_R$.
1465: By Proposition~\ref{76} $G_D^\circ$ is then also a connected finite quotient of $G/Z$ hence of $\Sl_m(\CC)\times G_R$, and GLCT holds for $D$ by Corollary~\ref{18}.
1466: \end{proof}
1467: 
1468: 
1469: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1470: \section{Classification in small dimensions}\label{48}
1471: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1472: 
1473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1474: \subsection{Structure of logarithmic vector fields}\label{49}
1475: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1476: 
1477: Let $\delta, \xi\in\Der$ and let $g\in\O$. 
1478: To emphasise the action of $\delta$ on $\O$ and on $\Der$, in place of $dg(\delta)$ we write $\delta(g)$, and in place of $[\delta,\xi]$ we write $\delta(\xi)=\ad_\delta(\xi)$.
1479: The degree $k$ parts of $\Ga(\CC^n,\Der)$ and $\Ga(\CC^n,\Der(-\log D))$ with respect to $\deg(x_i)=1=-\deg(\p_i)$ will be denoted by $\Ga(\CC^n,\Der)_k$ and $\Ga(\CC^n,\Der(-\log D))_k$ respectively.
1480: For $\delta\in\Ga(\CC^n,\Der)_0$, we write $\delta_S$ for its semisimple part and $\delta_N$ for its nilpotent part.
1481: 
1482: Let $D\subseteq\CC^n$ be a linear free divisor defined by the homogeneous polynomial $\Delta=\det((\delta_i(x_j))_{i,j})\in\CC[x]$ of degree $n$ as in Lemma~\ref{28} where $\delta=\delta_1,\dots,\delta_n$ is a global degree $0$ basis of $\Der(-\log D)$.
1483: Then $\delta_i(\Delta)\in\CC\cdot\Delta$ and there is the standard Euler vector field $\chi=\sum_ix_i\p_i\in\langle\delta_1,\dots,\delta_n\rangle_{\CC}$.
1484: Since $\chi(\Delta)/\Delta=n\ne0$, we can assume that $\delta_1=\chi$ and $\delta_i(\Delta)=0$ for $i=2,\dots,n$.
1485: So $\delta_2,\dots,\delta_n$ is a global degree $0$ basis of the annihilator $\Der(-\log\Delta)$ of $\Delta$ which is a direct factor of $\Der(-\log D)$.
1486: 
1487: Since $\chi$ vanishes only at the origin, the origin of the affine coordinate system $x=x_1,\dots,x_n$ is uniquely determined.
1488: A coordinate change between two degree $0$ bases of $\Der(-\log D)$ can always be chosen linear. 
1489: Among all possible linear coordinate changes, let $s+1$ be the maximal number of linearly independent diagonal logarithmic vector fields.
1490: 
1491: \begin{theo}\label{51}
1492: There exists a global degree $0$ basis $\chi$, $\sigma_1,\dots,\sigma_s$, $\nu_1,\dots,\nu_{n-s-1}$ 
1493: of $\Der(-\log D)$ such that
1494: \begin{enumerate}
1495: \item\label{51a} $\chi(\sigma_i)=0$ and $\chi(\nu_j)=0$,
1496: \item\label{51b} the $\sigma_i$ are simultaneously diagonalizable with eigenvalues in $\QQ$ and $\sigma_i(\Delta)=0$,
1497: \item\label{51c} the $\nu_j$ are nilpotent and $\nu_j(\Delta)=0$,
1498: \item\label{51d} $\sigma_i(\nu_j)\in\QQ\cdot\nu_j$ and 
1499: $\sum_j\sigma_i(\nu_j)/\nu_j+\tr(\sigma_i)=0$.
1500: \item\label{51e} If $\delta\in\Ga(\CC^n,\Der(-\log D))_0$ with $\sigma_i(\delta)=0$ for $i=1,\dots,s$ then $\delta_S\in\ideal{\sigma_1,\dots,\sigma_s}_\CC$.
1501: \end{enumerate}
1502: Moreover, $s\ge 1$ and if $s=n-1$ then $\Delta=x_1\cdots x_n$ defines a normal crossing divisor.
1503: \end{theo}
1504: 
1505: \begin{proof}
1506: It is easy to check that the formal coordinate changes used in \cite{GS06} reduce to \emph{linear} coordinate changes in the case of \emph{linear} free divisors.
1507: Thus \eqref{51a}-\eqref{51c}, \eqref{51e}, and the first part of \eqref{51d} follow from \cite[Thm.~5.4]{GS06}. 
1508: 
1509: For the second part of \eqref{51d}, we set $\delta_1,\dots,\delta_n=\chi,\sigma_1,\dots,\sigma_s,\nu_1\dots,\nu_{n-s-1}$ and rewrite $\Delta$ as
1510: \beq\label{27}
1511: \Delta=\sum_{\alpha\in S_n}\sgn(\alpha)\cdot\delta_1(x_{\alpha_1})\cdots\delta_n(x_{\alpha_n}).
1512: \eeq
1513: Let us choose coordinates in which all $\sigma _i$ are diagonal: $\sigma_i=\sum_j w_{i,j}x_j\p_j$. 
1514: The equation $\sigma_i(\Delta)=0$ means that $\Delta$ is weighted homogeneous of degree zero when we assign to the variable $x_j$ the weight $w_{i,j}=\sigma_i(x_j)/x_j$.
1515: The weighted degree of $\delta_j(x_k)$ is then $\sigma_i(\delta _j )/\delta_j+w_{i,k}$. 
1516: This implies the second part of \eqref{51d}, since each term in the sum \eqref{27} has the same weighted degree $\sum_j\big(\sigma_i(\delta_j)/\delta_j+w_{i \alpha_j}\big)=\sum_j\sigma_i(\delta_j)/\delta_j+\tr(\sigma_i)$.
1517: 
1518: Now assume that $s=0$.
1519: Then the vector space generated by the $\nu _i$ is entirely made of nilpotent elements and we can apply Engel's Theorem \cite[I.4]{Ser87}, and $\nu_1,\dots,\nu_{n-1}$ can be chosen upper triangular.
1520: But then $\Delta$ is clearly divisible by the square of the first variable $x_1$ and hence not reduced.
1521: So $s=0$ is impossible.
1522: 
1523: If $s=n-1$, then $\Delta$ must be a monomial and hence $\Delta=x_1\cdots x_n$ defines a normal crossing divisor.
1524: \end{proof}
1525: 
1526: \begin{rema}
1527: In Theorem~\ref{51}, one can perform the Gauss algorithm on the diagonals of $\sigma_1,\dots,\sigma_s$.
1528: Then $\sigma_i\equiv x_i\p_i\mod\sum_{j=s+1}^n\CC\cdot x_j\p_j$.
1529: \end{rema}
1530: 
1531: We shall frequently use the following simple fact.
1532: 
1533: \begin{lemm}\label{52}
1534: Let $\sigma=\sum_{i}w_ix_i\p_i$.
1535: Then $x_i\p_j$ is an eigenvector of $\ad_\sigma$ for the eigenvalue 
1536: $w_i-w_j$.
1537: \end{lemm}
1538: 
1539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1540: \subsection{The case $s=n-2$}\label{53}
1541: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1542: 
1543: \begin{lemm}\label{54}
1544: Let $s=n-2$ in the situation of Theorem~\ref{51}.
1545: Then $-\sigma_k(\nu_1)/\nu_1=\tr(\sigma_k)\ne0$ for some $k$ and $\nu_1$ has, after normalization, two entries equal to $1$ and all other entries equal to $0$.
1546: \end{lemm}
1547: 
1548: \begin{proof}
1549: If $s=n-2$ then for any $\sigma \in \{\sigma_1,\dots,\sigma_{n-2}\}$, $\sigma(\nu_1)/\nu_1+\tr(\sigma)=0$.
1550: Hence, a monomial $x_i\p_j$ in $\nu_1$ gives a relation $w_i-w_j+\sum_kw_k=0$ on the diagonal entries $w_1,\dots,w_n$ of $\sigma$.
1551: Since $3$ of these relations with $i\ne j$ and also $\sigma_1,\dots,\sigma_{n-2}$ are linearly independent, $\nu_1$ can have at most $2$ nonzero nondiagonal entries.
1552: If $\sigma_k(\nu_1)/\nu_1\ne0$ for some $k$ then $\nu_1$ is strictly triangular with at most $2$ nonzero entries after ordering the diagonal of $\sigma_k$.
1553: If $\nu_1$ has only one nonzero entry then $\Delta$ is divisible by the square of a variable, a contradiction.
1554: Both nonzero entries of $\nu_1$ can be normalized to $1$.
1555: If $\sigma_k(\nu_1)/\nu_1=0$ for all $k$ then the nonzero entries of $\nu_1$ are in a $2$-dimensional simultaneous eigenspace of $\chi,\sigma_1,\dots,\sigma_{n-2}$.
1556: Otherwise, there are $3$ linearly independent relations $w_{i_1}=w_{j_1}$, $w_{i_2}=w_{j_2}$, $\sum_kw_k=0$ on the diagonal entries of $\sigma_1,\dots,\sigma_{n-2}$, a contradiction to the linear independence of these vector fields.
1557: But then $\nu_1$ has only one nonzero entry after a linear coordinate change, a contradiction as before.
1558: \end{proof}
1559: 
1560: To simplify the notation, we shall write $\equiv$ for equivalence modulo $\CC^*$.
1561: By Lemma~\ref{54}, we may assume that $\nu_1=x_k\p_1+x_l\p_2$ where $1\ne k\ne l\ne 2$.
1562: Then 
1563: \[
1564: \Delta=
1565: \begin{vmatrix}
1566: x_1 & x_2 & x_3 & \cdots & x_n \\
1567: a_{2,1}x_1 & a_{2,2}x_2 & a_{2,3}x_3 & \cdots & a_{2,n}x_n \\
1568: \vdots & \vdots & \vdots && \vdots \\
1569: a_{n-1,1}x_1 & a_{n-1,2}x_2 & a_{n-1,3}x_3 & \cdots & a_{1,n}x_n \\
1570: x_k & x_l & 0 & \cdots & 0
1571: \end{vmatrix}
1572: \equiv(x_2x_k-x_1x_l)x_3\cdots x_n.
1573: \]
1574: 
1575: As $\Delta$ is reduced, there are, up to coordinate changes, only two non-normal-crossing cases:
1576: 
1577: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1578: \subsubsection{$k=2$, $l=3$}\label{55}
1579: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1580: 
1581: Then $\Delta$ comes from the linear free divisor of Example~\ref{40} in dimension $3$:
1582: \[
1583: \Delta=(x_2^2-x_1x_3)x_3\cdots x_n\equiv
1584: \begin{vmatrix}
1585: x_1 & x_2 & x_3 \\
1586: 4x_1 & x_2 & -2x_3 \\
1587: 2x_2 & x_3 & 0
1588: \end{vmatrix}
1589: \cdot x_4\cdots x_n.
1590: \]
1591: 
1592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1593: \subsubsection{$k=3$, $l=4$}
1594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1595: 
1596: Then $\Delta$ comes from a linear free divisor in dimension $4$:
1597: \[
1598: \Delta=(x_2x_3-x_1x_4)x_3\cdots x_n\equiv
1599: \begin{vmatrix}
1600: x_1 & x_2 & x_3 & x_4 \\
1601: x_1 & 2x_2 & -x_3 & 0 \\
1602: 2x_1 & x_2 & 0 & -x_4 \\
1603: x_3 & x_4 & 0 & 0
1604: \end{vmatrix}
1605: \cdot x_5\cdots x_n.
1606: \]
1607: 
1608: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1609: \subsection{Classification up to dimension $4$}\setcounter{secnumdepth}{5}
1610: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1611: 
1612: We consider the situation of Theorem~\ref{51} and abbreviate $x,y,z,w=x_1,x_2,x_3,x_4$.
1613: By the results of Section~\ref{53}, we may assume that $s=1$ and $n=4$.
1614: Let us first assume that $\Ga(\CC^n,\Der(-\log D))_0$ is a nonsolvable Lie algebra and hence $\langle\sigma_1,\nu_1,\nu_2\rangle=\sl_2$.
1615: 
1616: Recall that by \cite[IV.4]{Ser87}, $\CC^4$ is a direct sum of irreducible $\sl_2$-modules $W_m$ of dimension $m+1$ and that $W_m$ is represented in a basis $e_0,\dots,e_m$ by
1617: \[
1618: \sigma_1(e_i)=(-m+2i)e_i,\quad
1619: \nu_1(e_i)=(i+1)e_{i+1},\quad
1620: \nu_2(e_i)=(m-i+1)e_{i-1}.
1621: \]
1622: So there are 3 types of $\sl_2$-representations.
1623: The first two cases are $\CC^4=W_2\oplus W_0$ and $\CC^4=W_1\oplus W_1$, which lead to a zero and a nonreduced determinant of the form $\Delta\equiv(xw-yz)^2$ respectively.
1624: But $W_3$ gives the nontrivial linear free divisor
1625: \[
1626: \Delta=\begin{vmatrix}
1627: x&y&z&w\\
1628: -3x&-y&z&3w\\
1629: y&2z&3w&0\\
1630: 0&3x&2y&z
1631: \end{vmatrix}
1632: \equiv y^2z^2-4xz^3-4y^3w+18xyzw-27x^2w^2
1633: \]
1634: isomorphic to the discriminant in the space of binary cubics described in Example~\ref{1}.\eqref{1b}.
1635: 
1636: Now, assume that $\Ga(\CC^n,\Der(-\log D))_0$ is a solvable Lie algebra.
1637: Then, by Lie's Theorem \cite[I.7]{Ser87}, $\nu_1$ and $\nu_2$ can be chosen triangular and also $[\nu_1,\nu_2]$ is triangular.
1638: Hence, $[\nu_1,\nu_2]\in\langle\nu_1,\nu_2\rangle$ and even $[\nu_1,\nu_2]=0$ by nilpotency of $\ad_{\nu_1}$ and $\ad_{\nu_2}$.
1639: 
1640: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1641: \begin{annalif}
1642: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1643: 
1644: We set
1645: \[
1646: \sigma_1=ax\p_x+by\p_y+cz\p_z+dw\p_w
1647: \]
1648: and start a case by case discussion with respect to the cardinality of $\{a,b,c,d\}$. 
1649: In the following we shall omit the details that can be found in the preprint version \cite{GMNS06} of this article.
1650: 
1651: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1652: \subsubsection{$2\leq |\{a,b,c,d\}|\leq 3$}
1653: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1654: 
1655: In each case we refine to subcases depending on whether $\sigma_1(\nu_i)/\nu_i$, $i=1,2$, is zero or not.
1656: All these subcases lead to $\Delta=0$ or $\Delta$ being divisible by a square of a variable, in contradiction to our assumptions. 
1657: 
1658: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1659: \subsubsection{$|\{a,b,c,d\}|=4$}
1660: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1661: 
1662: Since $\nu_1$ and $\nu_2$ are $\sigma_1$-homogeneous and might be chosen triangular, $\sigma(\nu_1)/\nu_1\ne0\ne\sigma(\nu_2)/\nu_2$ and hence $\nu_1$ and $\nu_2$ have at most $3$ nonzero entries by Lemma~\ref{52}.
1663: Using, if necessary, permutations of the basis vectors, we have only to consider the following two cases: 
1664: 
1665: \paragraph{$\nu_1$ has one nonzero term or $\nu_1$ and $\nu_2$ have at most two nonzero terms.}
1666: 
1667: In both cases, a detailed discussion leads to the contradiction that $\Delta$ is divisible by a square of a variable.
1668: 
1669: \paragraph{$\nu _1 $ has three nonzero terms.}
1670: 
1671: This turns out to be the only case that leads to a linear free divisor with solvable Lie algebra and $s=1$.
1672: We may assume that
1673: \[
1674: \nu_1=
1675: \begin{pmatrix}
1676: 0 & 1 & 0 & 0 \\
1677: 0 & 0 & 1 & 0 \\
1678: 0 & 0 & 0 & 1 \\
1679: 0 & 0 & 0 & 0 
1680: \end{pmatrix}
1681: \]
1682: which implies also that $(a,b,c,d)=a\cdot(1,1,1,1)+(0,\lambda,2\lambda,
1683: 3\lambda)$ for some $0\neq\lambda\in\QQ$.
1684: The relation $[\nu_1,\nu_2]=0$ then implies that 
1685: \[
1686: \nu_2=p\cdot\nu_1+
1687: \begin{pmatrix}
1688: 0 & 0 & q & r \\
1689: 0 & 0 & 0 & q \\
1690: 0 & 0 & 0 & 0 \\
1691: 0 & 0 & 0 & 0 
1692: \end{pmatrix}.
1693: \] 
1694: So using the $\sigma_1$-homogeneity of $\nu_2$ the only case which was not yet considered is
1695: \[
1696: \nu_2= 
1697: \begin{pmatrix}
1698: 0 & 0 & 1 & 0 \\
1699: 0 & 0 & 0 & 1 \\
1700: 0 & 0 & 0 & 0 \\
1701: 0 & 0 & 0 & 0
1702: \end{pmatrix},\quad
1703: \Delta\equiv
1704: \begin{vmatrix}
1705: x & y & z & w \\
1706: 0 & \lambda y & 2\lambda z & 3\lambda w \\
1707: 0 & x & y & z \\
1708: 0 & 0 & x & y
1709: \end{vmatrix}
1710: \equiv x(y^3-3xyz+3x^2w).
1711: \] 
1712: The trace equation in Theorem~\ref{51}.\ref{51d} for $\sigma_1$ reads $-\lambda-2\lambda+4a+6\lambda=0$ or $a=-\frac{3}{4}\lambda $.
1713: Setting $\lambda=4$, we obtain $\sigma_1=-3x\p_x+y\p_y+5z\p_z+9w\p_w$.
1714: 
1715: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1716: \end{annalif}
1717: \begin{arxiv}
1718: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1719: 
1720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1721: \subsubsection{$\sigma_1=ax\p_x+ay\p_y+az\p_z+bw\p_w$, $a\ne b$}
1722: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1723: 
1724: Then $f=0$ if $\nu_1$ or $\nu_2$ has only a $\p_w$-component.
1725: We shall tacitly omit this case in the following.
1726: 
1727: \paragraph{$\sigma_1(\nu_1)/\nu_1=0=\sigma_1(\nu_2)/\nu_2$.}
1728: 
1729: Then 
1730: \[
1731: \nu_1=
1732: \begin{pmatrix}
1733: 0 & 1 & 0 & 0 \\
1734: 0 & 0 & \epsilon & 0 \\
1735: 0 & 0 & 0 & 0 \\
1736: 0 & 0 & 0 & 0 
1737: \end{pmatrix},\quad
1738: \nu_2=
1739: \begin{pmatrix}
1740: 0 & * & * & 0 \\
1741: 0 & 0 & * & 0 \\
1742: 0 & \overline\epsilon & 0 & 0 \\
1743: 0 & 0 & 0 & 0 
1744: \end{pmatrix},\quad
1745: f\equiv
1746: \begin{vmatrix}
1747: x & y & z & 0 \\
1748: 0 & 0 & 0 & w \\
1749: 0 & x & \epsilon y & 0 \\
1750: 0 & \overline\epsilon z & * & 0 
1751: \end{vmatrix}
1752: \]
1753: where $\epsilon\in\{0,1\}$ and $\overline\epsilon=1-\epsilon$.
1754: Hence, $f$ is divisible by $x^2$ and not reduced.
1755: 
1756: \paragraph{$\sigma_1(\nu_1)/\nu_1=0\ne\sigma_1(\nu_2)/\nu_2$.}
1757: 
1758: Then, after a linear coordinate change in $x,y,z$, we have
1759: \[
1760: \nu_2=
1761: \begin{pmatrix}
1762: 0 & 0 & 0 & 0 \\
1763: 0 & 0 & 0 & 0 \\
1764: 0 & 0 & 0 & 0 \\
1765: 1 & 0 & 0 & 0 
1766: \end{pmatrix},\quad
1767: f\equiv
1768: \begin{vmatrix}
1769: x & y & z & 0 \\
1770: 0 & 0 & 0 & w \\
1771: * & * & * & * \\
1772: w & 0 & 0 & 0
1773: \end{vmatrix}.
1774: \]
1775: Hence, $f$ is divisible by $w^2$ and not reduced.
1776: 
1777: \paragraph{$\sigma_1(\nu_1)/\nu_1\ne0\ne\sigma_1(\nu_2)/\nu_2$.}
1778: 
1779: Then 
1780: \[
1781: \nu_1=
1782: \begin{pmatrix}
1783: 0 & 0 & 0 & 0 \\
1784: 0 & 0 & 0 & 0 \\
1785: 0 & 0 & 0 & 0 \\
1786: 1 & 0 & 0 & 0 
1787: \end{pmatrix},\quad
1788: \nu_2=
1789: \begin{pmatrix}
1790: 0 & 0 & 0 & 0 \\
1791: 0 & 0 & 0 & 0 \\
1792: 0 & 0 & 0 & 0 \\
1793: r & s & t & 0 
1794: \end{pmatrix}
1795: \]
1796: and hence $f$ is divisible by $w^3$ and not reduced.
1797: 
1798: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1799: \subsubsection{$\sigma_1=ax\p_x+ay\p_y+bz\p_z+bw\p_w$, $a\ne b$}
1800: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1801: 
1802: Then $f=0$ if $\nu_1$ and $\nu_2$ have only a $\p_z$- and $\p_w$-component.
1803: We shall tacitly omit this case in the following.
1804: 
1805: \paragraph{$\sigma_1(\nu_1)/\nu_1=0=\sigma_1(\nu_2)/\nu_2$.}
1806: 
1807: Then 
1808: \[
1809: \nu_1=
1810: \begin{pmatrix}
1811: 0 & 1 & 0 & 0 \\
1812: 0 & 0 & 0 & 0 \\
1813: 0 & 0 & 0 & 0 \\
1814: 0 & 0 & 0 & 0 
1815: \end{pmatrix},\quad
1816: \nu_2=
1817: \begin{pmatrix}
1818: 0 & 0 & 0 & 0 \\
1819: 0 & 0 & 0 & 0 \\
1820: 0 & 0 & 0 & 1 \\
1821: 0 & 0 & 0 & 0 
1822: \end{pmatrix},\quad
1823: f\equiv
1824: \begin{vmatrix}
1825: x & y & 0 & 0 \\
1826: 0 & 0 & z & w \\
1827: 0 & x & 0 & 0 \\
1828: 0 & 0 & 0 & z
1829: \end{vmatrix}\equiv x^2z^2.
1830: \]
1831: 
1832: \paragraph{$\sigma_1(\nu_1)/\nu_1=0\ne\sigma_1(\nu_2)/\nu_2$.}
1833: 
1834: Then 
1835: \[
1836: \nu_1=
1837: \begin{pmatrix}
1838: 0 & 1 & 0 & 0 \\
1839: 0 & 0 & 0 & 0 \\
1840: 0 & 0 & 0 & \epsilon \\
1841: 0 & 0 & 0 & 0 
1842: \end{pmatrix},\quad
1843: \nu_2=
1844: \begin{pmatrix}
1845: 0 & 0 & * & * \\
1846: 0 & 0 & * & * \\
1847: 0 & 0 & 0 & 0 \\
1848: 0 & 0 & 0 & 0 
1849: \end{pmatrix},\quad
1850: f\equiv
1851: \begin{vmatrix}
1852: x & y & 0 & 0 \\
1853: 0 & 0 & z & w \\
1854: 0 & x & 0 & \epsilon z \\
1855: 0 & 0 & * & *
1856: \end{vmatrix}
1857: \]
1858: where $\epsilon\in\{0,1\}$. 
1859: Hence, $f$ is divisible by $x^2$ (or $z^2$ for transposed $\nu_2$) and 
1860: not reduced.
1861: 
1862: \paragraph{$\sigma_1(\nu_1)/\nu_1\ne0\ne\sigma_1(\nu_2)/\nu_2$.}
1863: 
1864: Then 
1865: \[
1866: \nu_1=
1867: \begin{pmatrix}
1868: 0 & 0 & 1 & 0 \\
1869: 0 & 0 & 0 & 0 \\
1870: 0 & 0 & 0 & 0 \\
1871: 0 & 0 & 0 & 0 
1872: \end{pmatrix},\quad
1873: \nu_2=
1874: \begin{pmatrix}
1875: 0 & 0 & 0 & 0 \\
1876: 0 & 0 & 0 & 0 \\
1877: 0 & 0 & 0 & 0 \\
1878: 0 & 1 & 0 & 0 
1879: \end{pmatrix},\quad
1880: f\equiv
1881: \begin{vmatrix}
1882: x & y & 0 & 0 \\
1883: 0 & 0 & z & w \\
1884: 0 & 0 & x & 0 \\
1885: 0 & w & 0 & 0
1886: \end{vmatrix}\equiv x^2w^2.
1887: \]
1888: 
1889: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1890: \subsubsection{$\sigma_1=ax\p_x+ay\p_y+bz\p_z+cw\p_w$, $|\{a,b,c\}|=3$}
1891: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1892: 
1893: Then $f=0$ if $\nu_1$ and $\nu_2$ have only a $\p_z$- and $\p_w$-component.
1894: We shall tacitly omit this case in the following.
1895: 
1896: \paragraph{$\sigma_1(\nu_1)/\nu_1=0=\sigma_1(\nu_2)/\nu_2$.}
1897: 
1898: Then $\nu_1$ and $\nu_2$ are linearly dependent.
1899: 
1900: \paragraph{$\sigma_1(\nu_1)/\nu_1=0\ne\sigma_1(\nu_2)/\nu_2$.}
1901: 
1902: Then 
1903: \[
1904: \nu_1=
1905: \begin{pmatrix}
1906: 0 & 1 & 0 & 0 \\
1907: 0 & 0 & 0 & 0 \\
1908: 0 & 0 & 0 & 0 \\
1909: 0 & 0 & 0 & 0 
1910: \end{pmatrix},\quad
1911: \nu_2=
1912: \begin{pmatrix}
1913: 0 & 0 & * & * \\
1914: 0 & 0 & 0 & 0 \\
1915: 0 & * & 0 & * \\
1916: 0 & * & * & 0 
1917: \end{pmatrix},\quad
1918: f\equiv
1919: \begin{vmatrix}
1920: x & y & z & w \\
1921: 0 & 0 & bz & cw \\
1922: 0 & x & 0 & 0 \\
1923: 0 & * & * & *
1924: \end{vmatrix}.
1925: \]
1926: Hence, $f$ is divisible by $x^2$ and not reduced.
1927: 
1928: \paragraph{$\sigma_1(\nu_1)/\nu_1\ne0\ne\sigma_1(\nu_2)/\nu_2$.}
1929: 
1930: Then there are 3 double cases:
1931: \begin{enumerate}
1932: 
1933: \item
1934: \[
1935: \nu_1=
1936: \begin{pmatrix}
1937: 0 & 0 & 1 & 0 \\
1938: 0 & 0 & 0 & 0 \\
1939: 0 & 0 & 0 & 0 \\
1940: 0 & 0 & 0 & 0 
1941: \end{pmatrix},\quad
1942: \nu_2=
1943: \begin{pmatrix}
1944: 0 & 0 & * & * \\
1945: 0 & 0 & * & * \\
1946: 0 & 0 & 0 & 0 \\
1947: 0 & * & * & 0 
1948: \end{pmatrix},\quad
1949: f\equiv
1950: \begin{vmatrix}
1951: x & y & z & w \\
1952: 0 & 0 & bz & cw \\
1953: 0 & 0 & x & 0 \\
1954: 0 & * & * & *
1955: \end{vmatrix}.
1956: \]
1957: Hence, $f$ is divisible by $x^2$ (or $z^2$ for transposed $\nu_1$ and 
1958: $\nu_2$) and not reduced.
1959: 
1960: \item
1961: \[
1962: \nu_1=
1963: \begin{pmatrix}
1964: 0 & 0 & 0 & 0 \\
1965: 0 & 0 & 0 & 0 \\
1966: 0 & 0 & 0 & 1 \\
1967: 0 & 0 & 0 & 0 
1968: \end{pmatrix},\quad
1969: \nu_2=
1970: \begin{pmatrix}
1971: 0 & 0 & 0 & * \\
1972: 0 & 0 & 0 & * \\
1973: * & * & 0 & 0 \\
1974: 0 & 0 & 0 & 0 
1975: \end{pmatrix},\quad
1976: f\equiv
1977: \begin{vmatrix}
1978: x & y & z & w \\
1979: 0 & 0 & bz & cw \\
1980: 0 & 0 & 0 & z \\
1981: * & * & 0 & *
1982: \end{vmatrix}.
1983: \]
1984: Hence, $f$ is divisible by $z^2$ (or $w^2$ for transposed $\nu_1$ and 
1985: $\nu_2$) and not reduced.
1986: 
1987: \item
1988: \[
1989: \nu_1=
1990: \begin{pmatrix}
1991: 0 & 0 & 1 & 0 \\
1992: 0 & 0 & 0 & 0 \\
1993: 0 & 0 & 0 & 1 \\
1994: 0 & 0 & 0 & 0 
1995: \end{pmatrix},\quad
1996: \nu_2=
1997: \begin{pmatrix}
1998: 0 & 0 & 0 & * \\
1999: 0 & 0 & 0 & * \\
2000: 0 & 0 & 0 & 0 \\
2001: 0 & 0 & 0 & 0 
2002: \end{pmatrix},\quad
2003: f\equiv
2004: \begin{vmatrix}
2005: x & y & z & w \\
2006: 0 & 0 & bz & cw \\
2007: 0 & 0 & x & z \\
2008: 0 & 0 & 0 & *
2009: \end{vmatrix}=0
2010: \]
2011: or $f$ is divisible by $w^2$ for transposed $\nu_1$ and $\nu_2$ and not 
2012: reduced.
2013: 
2014: \end{enumerate}
2015: 
2016: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2017: \subsubsection{$\sigma_1=ax\p_x+by\p_y+cz\p_z+dw\p_w$, $|\{a,b,c,d\}|=4$}
2018: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2019: 
2020: Since $\nu_1$ and $\nu_2$ are $\sigma_1$-homogeneous and might be chosen 
2021: triangular, $\sigma(\nu_1)/\nu_1\ne0\ne\sigma(\nu_2)/\nu_2$ and hence 
2022: $\nu_1$ and $\nu_2$ have at most $3$ non-zero entries by Lemma~\ref{52}.
2023: Using if necessary permutations of the basis vectors, we have only to 
2024: consider the following cases: 
2025: 
2026: \paragraph{$\nu _1$ has one non-zero term.}
2027: 
2028: We may assume that 
2029: \[
2030: \nu_1=
2031: \begin{pmatrix}
2032: 0 & 1 & 0 & 0 \\
2033: 0 & 0 & 0 & 0 \\
2034: 0 & 0 & 0 & 0 \\
2035: 0 & 0 & 0 & 0 
2036: \end{pmatrix}.
2037: \] 
2038: Since $[\nu_1,\nu_2]=0$, we find that the first column of $\nu_2$ is 
2039: zero and hence
2040: \[
2041: f\equiv
2042: \begin{vmatrix}
2043: x & y & z & w \\
2044: ax & by & cz & dw \\
2045: 0 & x & 0 & 0 \\
2046: 0 & * & * & *
2047: \end{vmatrix}
2048: \]
2049: is non-reduced as a multiple of $x^2$. 
2050: In what follows, we may also exclude all cases where $\nu_2$ has only 
2051: one term by exchanging the roles of $\nu_1$ and $\nu_2$. 
2052: 
2053: \paragraph{$\nu_1$ has two non-zero terms.}
2054: We have two cases:
2055: \[
2056: \nu_1=
2057: \begin{pmatrix}
2058: 0 & 1 & 0 & 0 \\
2059: 0 & 0 & 1 & 0 \\
2060: 0 & 0 & 0 & 0 \\
2061: 0 & 0 & 0 & 0 
2062: \end{pmatrix}\quad
2063: \text{ or }\quad
2064: \nu_1=
2065: \begin{pmatrix}
2066: 0 & 1 & 0 & 0 \\
2067: 0 & 0 & 0 & 0 \\
2068: 0 & 0 & 0 & 1 \\
2069: 0 & 0 & 0 & 0 
2070: \end{pmatrix}.
2071: \] 
2072: In the first case, since $[\nu_1,\nu_2]=0$, we obtain that $\nu _2$ has 
2073: the form
2074: \[
2075: \nu_2=p\cdot\nu_1+
2076: \begin{pmatrix}
2077: 0 & 0 & q & r \\
2078: 0 & 0 & 0 & 0 \\
2079: 0 & 0 & 0 & 0 \\
2080: 0 & 0 & s & 0 
2081: \end{pmatrix}.
2082: \] 
2083: Since $a-b=b-c$ because of the homogeneity of $\nu_1$, the $\sigma_1$-degrees 
2084: corresponding to $r,s$ are $a-d\neq d-c$.
2085: Then, because of the homogeneity of $\nu_2$, we are in the situation 
2086: where $\nu _2$ has only one term. 
2087: In the second case, we obtain 
2088: \[
2089: \nu_2=p\cdot\nu_1+
2090: \begin{pmatrix}
2091: 0 & 0 & u & v \\
2092: 0 & 0 & 0 & u \\
2093: r & s & 0 & q \\
2094: 0 & r & 0 & 0 
2095: \end{pmatrix}.
2096: \] 
2097: If $u=1$ and $v=r=s=q=0$ then 
2098: \[
2099: \nu_2=p\cdot\nu_1+
2100: \begin{pmatrix}
2101: 0 & 0 & 1 & 0 \\
2102: 0 & 0 & 0 & 1 \\
2103: 0 & 0 & 0 & 0 \\
2104: 0 & 0 & 0 & 0 
2105: \end{pmatrix},\quad
2106: f\equiv
2107: \begin{vmatrix}
2108: x & y & z & w \\
2109: ax & by & cz & dw \\
2110: 0 & x & 0 & z \\
2111: 0 & 0 & x & y
2112: \end{vmatrix}.
2113: \]
2114: So $f$ is non-reduced as a multiple of $x^2$.  
2115: Similarly, if $r=1$ and $u=v=s=q=0$ then $f$ is a multiple of $z^2$.
2116: In the other cases,  either $\nu _2$ has only one non-zero term or 
2117: three non-zero terms ($u=s=1$).
2118: This latter case we shall study now.
2119: 
2120: \paragraph{$\nu _1 $ has three non-zero terms.}
2121: We may assume that
2122: \[
2123: \nu_1=
2124: \begin{pmatrix}
2125: 0 & 1 & 0 & 0 \\
2126: 0 & 0 & 1 & 0 \\
2127: 0 & 0 & 0 & 1 \\
2128: 0 & 0 & 0 & 0 
2129: \end{pmatrix}
2130: \]
2131: which implies also that $(a,b,c,d)=a\cdot(1,1,1,1)+(0,\lambda,2\lambda,
2132: 3\lambda)$ for some $0\neq\lambda\in\QQ$.
2133: The relation $[\nu_1,\nu_2]=0$ then implies that 
2134: \[
2135: \nu_2=p\cdot\nu_1+
2136: \begin{pmatrix}
2137: 0 & 0 & q & r \\
2138: 0 & 0 & 0 & q \\
2139: 0 & 0 & 0 & 0 \\
2140: 0 & 0 & 0 & 0 
2141: \end{pmatrix}.
2142: \] 
2143: So the only remaining possibility not yet considered is
2144: \[
2145: \nu_2-p\nu _1= 
2146: \begin{pmatrix}
2147: 0 & 0 & 1 & 0 \\
2148: 0 & 0 & 0 & 1 \\
2149: 0 & 0 & 0 & 0 \\
2150: 0 & 0 & 0 & 0
2151: \end{pmatrix},\quad
2152: f\equiv
2153: \begin{vmatrix}
2154: x & y & z & w \\
2155: 0 & \lambda y & 2\lambda z & 3\lambda w \\
2156: 0 & x & y & z \\
2157: 0 & 0 & x & y
2158: \end{vmatrix}
2159: \equiv x(y^3-3xyz+3x^2w).
2160: \] 
2161: The equation $\sigma_1(\nu_1)/\nu_1+\sigma_1(\nu_2)/\nu_2+\tr(\sigma_1)=0$ 
2162: becomes $-\lambda-2\lambda+4a+6\lambda=0$ or $a=-\frac{3}{4}\lambda $.
2163: Setting $\lambda=4$, we obtain $\sigma_1=-3x\p_x+y\p_y+5z
2164: \p_z+9w\p_w$ and $\sigma_1(f)=0$.
2165: 
2166: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2167: \end{arxiv}
2168: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2169: 
2170: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2171: \subsection{Summary of the classification up to dimension 4}\label{56}
2172: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2173: 
2174: The following Table~\ref{88} summarizes our classification of linear free divisors up to dimension $4$.
2175: The matrices are interpreted row-wise as bases of $\Der(-\log D)$.
2176: 
2177: \begin{longtable}{|c|c|c|c|c|}
2178: \caption{Classification of linear free divisors up to dimension $4$}\label{88}\\
2179: \hline
2180: \rule{0pt}{11pt}$n$ & $\Delta$ & $\Der(-\log D)$ & $\gg_D$ & reductive?\\
2181: \hline
2182: \hline
2183: $1$ & $x$&$\begin{pmatrix}x\end{pmatrix}$ & $\CC$ & Yes \\
2184: \hline
2185: \hline
2186: $2$ & $xy$&$\begin{pmatrix}x&0\\0&y\end{pmatrix}$ & $\CC^2$ & Yes \\
2187: \hline
2188: \hline
2189: $3$ & $xyz$&$\begin{pmatrix}x&0&0\\0&y&0\\0&0&z\end{pmatrix}$ & $\CC^3$ & Yes \\
2190: \hline
2191: $3$& $(y^2+xz)z$&$\begin{pmatrix}x&y&z\\4x&y&-2z\\-2y&z&0\end{pmatrix}$ & $\bb_2$ & No \\
2192: \hline
2193: \hline
2194: $4$ & $xyzw$&$\begin{pmatrix}
2195: x&0&0&0\\0&y&0&0\\0&0&z&0\\0&0&0&w\end{pmatrix}$ & $\CC^4$ & Yes \\
2196: \hline
2197: $4$ & $(y^2+xz)zw$&$\begin{pmatrix}
2198: x&y&z&0\\4x&y&-2z&0\\-2y&z&0&0\\0&0&0&w\end{pmatrix}$ & $\CC\oplus\bb_2$ & No \\
2199: \hline
2200: $4$ & $(yz+xw)zw$&
2201: $\begin{pmatrix}
2202: x&0&0&-w\\0&y&0&w\\0&0&z&w\\z&-w&0&0
2203: \end{pmatrix}$ 
2204: & $\CC^2\oplus \gg_0$ & No \\
2205: \hline
2206: $4$ & $x(y^3-3xyz+3x^2w)$&
2207: $\begin{pmatrix}
2208: x&y&z&w\\0&y&2z&3w\\
2209: 0&x&y&z\\0&0&x&y
2210: \end{pmatrix}$ 
2211: & $\CC\oplus \gg$ & No \\
2212: \hline
2213: $4$ & \begin{minipage}{3cm}\begin{center}$y^2z^2-4xz^3-4y^3w+18xyzw-27w^2x^2$\end{center}\end{minipage}&
2214: $\begin{pmatrix}
2215: 3x&2y&z&0\\0&3x&2y&z\\y&2z&3w&0\\0&y&2z&3w
2216: \end{pmatrix}$ & $\gl_2(\CC)$ & Yes \\
2217: \hline
2218: \end{longtable}
2219: 
2220: The annihilators of $\Delta$ in the Lie algebras for $\Delta=(yz+xw)zw$ and $\Delta=x(y^3-3xyz+3x^2w)$ are described in \cite[Ch.~I, \S4]{Jac79}.
2221: The former is the direct sum of $\CC$ and the non-Abelian Lie algebra $\gg_0$ of dimension $2$, and the latter is the $3$-dimensional Lie algebra $\gg$ characterized as having $2$-dimensional Abelian derived algebra $\gg'$, on which the adjoint action of a basis vector outside $\gg'$ is semi-simple with eigenvalues $1$ and $2$. 
2222: Straightforward computations show that the two \emph{groups} $G_D^\circ$ are, respectively, the set of $4\times 4$ matrices of the form
2223: \[
2224: \begin{pmatrix}
2225: x^{-1}y^{-2}&0&z&0\\
2226: 0&x^{-2}y^{-1}&0&-x^{-1}yz\\
2227: 0&0&x&0\\
2228: 0&0&0&y
2229: \end{pmatrix}
2230: \quad
2231: \mbox{and}
2232: \quad
2233: \begin{pmatrix}
2234: x^{-3}&0&0&0\\
2235: y&x&0&0\\
2236: z&x^4y&x^5&0\\
2237: x^3yz-\frac{1}{3}x^6y^3&x^4z&x^8y&x^9
2238: \end{pmatrix}
2239: \] 
2240: with $x,y\in\CC^*$ and $z\in \CC$ in the first and $x\in\CC^*, y,z\in\CC$ in the second.
2241: 
2242: 
2243: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2244: \section{Strong Euler homogeneity and local quasihomogeneity}\label{5}
2245: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2246: 
2247: In this section, we investigate linear free divisors with respect to the properties of local quasihomogeneity and strong Euler homogeneity from Definitions~\ref{80} and \ref{81}.
2248: 
2249: The following reformulation of the definition of local quasihomogeneity is a direct consequence of the Poincar\'e--Dulac Theorem \cite[Ch.~3, \S 3.2]{ABGI88} and  Artin's Approximation Theorem \cite{Art68}.
2250: 
2251: \begin{theo}
2252: A divisor $D$ is locally quasihomogeneous if and only if, at any $p\in D$, there is an Euler vector field $\chi$ for $D$ at $p$ whose degree zero part $\chi_0$ has strictly positive eigenvalues.
2253: \end{theo}
2254: 
2255: We denote by $\D=\D_{\CC^n}$ the sheaf of germs of linear differential operators with holomorphic coefficients on $\CC^n$.
2256: It is naturally equipped with an increasing filtration $F$ of coherent $\O$-modules by the order of differential operators and we denote by $\sigma(P)$ the symbol of $P\in\D$ in $\gr_F\D$.
2257: Note that $\gr_F\D_p\cong\O_p[\p]=\CC\{x\}[\p]$ in a local coordinate system $x$ at $p$, where we identify $\sigma(\p_i)=\p_i$.The following property is closely related to local quasihomogeneity.
2258: 
2259: \begin{defi}
2260: A free divisor $D$ is called \emph{Koszul free} if, at any $p\in D$, there exists a basis $\delta_1,\dots,\delta_n$ of $\Der(-\log D)_p$ such that $\sigma(\delta_1),\dots,\sigma(\delta_n)$ is a regular sequence in $\gr_F\D_p$.
2261: \end{defi}
2262: 
2263: Koszul freeness can be interpreted geometrically in terms of the \emph{logarithmic stratification}, introduced by K.~Saito \cite{Sai80}, which is the partition of $D$ into the integral varieties of the distribution $\Der(-\log D)$. 
2264: As it is not always locally finite, the term ``stratification'' is a misnomer, but is generally used. 
2265: If $D_\alpha$ is a stratum of the logarithmic stratification and $p\in D_\alpha$ then $T_pD_\alpha=\Der(-\log D)(p)$.
2266: The graded ring $\gr_F\D_p\cong\O_p[\p]$ contains $\Der_p=\oplus_{i=1}^n\O_p\p_i$ and can be identified with the ring of functions on the cotangent space $T^*(\CC^n,p)$ of the germ $(\CC^n,p)$, polynomial on the fibers and analytic on the base.
2267: 
2268: \begin{defi}
2269: The \emph{logarithmic characteristic variety} $L_{\CC^n}(D)$ of $D$ is the variety in $T^*\CC^n$ defined by the image of $\Der(-\log D)$ in $\gr_F\D$.
2270: \end{defi}
2271: 
2272: Thus $D$ is Koszul free at $p$ if and only if $L_{\CC^n}(D)$ is purely $n$-dimensional \cite[1.8]{CN02}. 
2273: Moreover, by \cite[3.16]{Sai80}, $L_{\CC^n}(D)$ is the union, over all strata $D_\alpha\subseteq D$ in the logarithmic stratification of $D$, of the conormal bundle $T^*_{D_\alpha}\CC^n$ of $D_\alpha$, each of which is $n$-dimensional.
2274: This proves the following result.
2275: 
2276: \begin{theo}
2277: A free divisor $D$ is Koszul free if and only if the logarithmic stratification is locally finite. 
2278: \end{theo}
2279: 
2280: A locally quasihomogeneous free divisor is Koszul free \cite[4.3]{CN02}, and thus the failure of Koszul-freeness serves as a computable criterion for the failure of local quasihomogeneity.
2281: 
2282: As in Section \ref{49}, let $D\subseteq\CC^n$ be a linear free divisor and $\chi,\delta_2,\dots,\delta_n$ a global degree $0$ basis of $\Der(-\log D)$ with $\delta_i(\Delta)=0$ for $\Delta=\det(S)$ where
2283: \[
2284: S:=\begin{pmatrix}\chi(x_j)\\\delta_i(x_j)\end{pmatrix}_{i,j}
2285: \]
2286: The arguments which follow are valid as well for an arbitrary germ of a free divisor $D\subseteq(\CC^n,0)$ with a germ of an Euler vector field $\chi\in\Der_0$ at $0$. 
2287: 
2288: The following criterion gives a method to test strong Euler homogeneity algorithmically.
2289: The reduced variety $S_k$ defined by the $(k+1)\times(k+1)$-minors of the $n\times n$-matrix $S$ is the union of logarithmic strata of dimension at most $k$.
2290: In more invariant terms, $S_k$ is the variety of zeros of the $(n-k-1)$'st Fitting ideal of the $\O_{\CC^n}$-module $\Der/\Der(-\log D)$.
2291: Thus a free divisor $D$ is Koszul free if and only if $\dim S_k\leq k$ for all $k$. 
2292: Note that $S_n=\CC^n$, $S_{n-1}=D$, and $S_{n-2}=\Sing(D)$.
2293: For a linear free divisor, $S_0=\{0\}$ because of the presence of the Euler vector field. 
2294: Since $\dim\Sing(D)<\dim D$, it follows that linear free divisors are Koszul free in 
2295: dimension $n\le3$.
2296:  
2297: In order to characterize strong Euler homogeneity, we also consider the reduced variety $T_k\supseteq S_k$ defined by the $(k+1)\times(k+1)$-minors of the $(n-1)\times n$-matrix 
2298: \[
2299: T:=(\delta_i(x_j))_{i,j}.
2300: \]
2301: Again, this is the variety defined by a Fitting ideal, this time the $(n-k-2)$nd Fitting ideal of the module $\Der/\Der(-\log\Delta)$.
2302: Note that, by definition,
2303: \[
2304: S=\begin{pmatrix}x\\T\end{pmatrix}.
2305: \]
2306: 
2307: \begin{lemm}\label{57}
2308: $D$ is strongly Euler homogeneous if and only if $S_k=T_k$ for $0\le k\le n-2$.
2309: \end{lemm}
2310: 
2311: \begin{proof}
2312: A vector field $\delta\in\big(\mm_p\cdot\chi+\sum_i\O_p\cdot\delta_i\big)\cap\mm_p\cdot\Der_p$ is not an Euler vector field at $p\in D$ since $\delta(\Delta)\in\mm_p\cdot\Delta$, and indeed $\delta(u\cdot\Delta)\in\mm_p\cdot u\Delta$ for any unit $u$.
2313: Hence, an Euler vector field $\eta$ for $D$ at $p$ must be of the form $\eta=a_0\cdot\chi+\sum_ia_i\cdot\delta_i\in\mm_p\cdot\Der_p$ with $a_0(0)\neq 0$. 
2314: This means that $\chi(p)\in\sum_i\CC\cdot\delta_i(p)$, and the matrices $S$ and $T$ have equal rank at $p$. 
2315: Conversely if $\chi(p)=\sum \lambda_i\delta_i(p)$ then up to multiplication by a scalar, $\chi-\sum_i\lambda_i\delta_i$ is an Euler field for $D$ at $p$.
2316: \end{proof}
2317: 
2318: \begin{rema}
2319: The proof of Lemma~\ref{57} shows that the question of local 
2320: quasihomogeneity is much more complicated:
2321: The degree zero parts of Euler vector fields at a point $p\in D$ are the degree zero parts of vector fields $a_1\cdot\chi+\sum_ia_i\cdot\delta_i$ where $a_1,\dots,a_n$ are linear forms such that $a_1(p)\cdot\chi(p)+\sum_ia_i(p)\cdot\delta_i(p)=0$.
2322: \end{rema}
2323: 
2324: For $k=1,\dots,n$, let $M_k=(-1)^{k+1}\det (\delta_i(x_j))_{j\neq k}$.
2325: 
2326: \begin{lemm}\label{58}
2327: For $k=1,\dots,n$, $\p_k(\Delta)=n\cdot M_k$.
2328: In particular, $S_k=T_k$ for $k=n-2$.
2329: \end{lemm}
2330: 
2331: \begin{proof}
2332: Since
2333: \[
2334: S\begin{pmatrix}\p_1(\Delta)\\\vdots\\\p_n(\Delta)\end{pmatrix}=
2335: \begin{pmatrix}\chi\\\delta_2\\\vdots\\\delta_n\end{pmatrix}(\Delta)=
2336: \begin{pmatrix}n\cdot\Delta\\0\\\vdots\\0\end{pmatrix}
2337: \]
2338: we obtain, by canceling $\Delta$,
2339: \[
2340: \begin{pmatrix}\p_1(\Delta)\\\vdots\\\p_n(\Delta)\end{pmatrix}=\check 
2341: S\begin{pmatrix}n\\0\\\vdots\\0\end{pmatrix}=n\cdot\begin{pmatrix}M_1\\
2342: \vdots\\M_n\end{pmatrix}
2343: \]
2344: where $\check S$ denotes the cofactor matrix of $S$.
2345: \end{proof}
2346: 
2347: \begin{lemm}\label{59}
2348: $S_0=T_0$.
2349: \end{lemm}
2350: 
2351: \begin{proof}
2352: Assume that $T_0\ne S_0=\{0\}$.
2353: By homogeneity, $T_0$ contains the $x_n$-axis after an appropriate linear 
2354: coordinate change.
2355: Then $T$ is independent of $x_n$.
2356: Writing $x'=x_1,\dots,x_{n-1}$, we have that $\Delta=g+x_n\cdot\Delta'$ where $g$ and $\Delta':=M_n$ depend only on $x'$.
2357: Since $\Delta$ does not depend on fewer variables, we must have $\Delta'\ne0$.
2358: For $i=2,\dots,n$, let $\delta_i':=\sum_{j=1}^{n-1}\delta_i(x_j)\p_j$ 
2359: be the projection of $\delta_i$ to the $\CC[x]$-module with basis 
2360: $\p':=\p_1,\dots,\p_{n-1}$.
2361: Then, for $i=2,\dots,n$, $\delta'_i(\Delta')=0$, as it is the coefficient of $x_n$ in $\delta_i(\Delta)=0$.
2362: Since the rank of the $\CC[x']$-annihilator of $\p_1(\Delta'),\dots,\p_{n-1}(\Delta')$ is strictly smaller than $n-1$, 
2363: there must be a relation 
2364: $\sum_{i=2}^na_i\delta'_i=0$ for some homogeneous polynomials 
2365: $a_i\in\CC[x']$.
2366: But since $\delta_2,\dots,\delta_n$ are independent over $\CC[x]$, $\sum_{i=2}^na_i\delta_i(x_n)\ne0$ and hence $0=\sum_{i=2}^na_i\delta_i(\Delta)=\bigl(\sum_{i=2}^na_i\delta_i(x_n)\bigr)\cdot\Delta'$, contradicting the fact that $\Delta'\ne0$.
2367: \end{proof}
2368: 
2369: \begin{lemm}\label{60}
2370: Let $D$ be strongly Euler homogeneous.
2371: Then $D$ is locally quasihomogeneous on the complement of $S_{n-3}$.
2372: In particular, $D$ is locally quasihomogeneous if $S_{n-3}=\{0\}$.
2373: \end{lemm}
2374: 
2375: \begin{proof}
2376: By \cite[3.5]{Sai80}, $(D,p)=(D',p')\times(\CC^{n-2},0)$ for $p\in S_{n-2}\ssm S_{n-3}$ where $(D',p')\subseteq(\CC^2,0)$ is strongly Euler homogeneous by \cite[3.2]{GS06}.
2377: As the germ of a curve, $(D',p')$ has an isolated singularity.
2378: Then $(D',p')$, and hence $(D,p)$, are quasihomogeneous, by Saito's theorem \cite{Sai71}.
2379: \end{proof}
2380: 
2381: \begin{theo}\label{61}
2382: Every linear free divisor in dimension $n\le 4$ is locally quasihomogeneous and hence LCT and GLCT hold.
2383: \end{theo}
2384: 
2385: \begin{proof}
2386: By Lemmas \ref{58} and \ref{59}, $S_1=T_1$ if $n=3$ and $S_0=T_0$. 
2387: If $n\le3$ then $D$ is strongly Euler homogeneous by Lemma~\ref{57} and so locally quasihomogeneous by Lemma~\ref{60}.
2388: 
2389: For $n=4$ analogous arguments yield $S_0=T_0=\{0\}$ and $S_2=T_2$.
2390: Now we use the classification in Subsection \ref{56} and a case by case study:
2391: In each case, one can verify that $S_1=T_1$ and construct an Euler vector field with positive eigenvalues in degree zero at each point of $S_1\smallsetminus S_0$.
2392: Again this is sufficient for local quasihomogeneity by Lemma \ref{60}.
2393: For $\Delta=(yz+xw)zw$, $S_1=\{xy=z=w=0\}$ and $2\chi-\sigma+\frac{x-\xi}{\xi}\sigma$, where $\sigma=2x\p_x+y\p_y-w\p_w$, is an Euler vector field at $(\xi,0,0,0)\not\in S_0$ with eigenvalues $2,1,2,3$ in degree zero.
2394: For $\Delta=x(y^3-3xyz+3x^2w)$, $S_1=\{x=y=z=0\}$ and $9\chi-\sigma+\frac{w-\om}{\om}\sigma$, where $\sigma=-3x\p_x+y\p_y+5z\p_z+9w\p_w$, is an Euler vector field at $(0,0,0,\om)\not\in S_0$ with eigenvalues $12,8,4,9$ in degree zero.
2395: The remaining cases are trivial.
2396: 
2397: By \cite{CMN96}, local quasihomogeneity implies that LCT and hence, by taking global sections, GLCT holds.
2398: \end{proof}
2399: 
2400: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2401: \subsection{Example~\ref{41} again}\label{62}
2402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2403: 
2404: In this subsection, we study the linear free divisor in Example~\ref{41} in detail and show that if $n>2$ it is not Koszul free and hence not locally quasihomogeneous.
2405: However we will see that it is strongly Euler homogeneous, like all other linear free divisors whose strong Euler homogeneity has been investigated.
2406: 
2407: Denote by $x_{i,j}$, $1\leq  i\leq n$, $1\leq  j\leq n+1$, the coordinates on the space of $n\times(n+1)$-matrices $M_{n,n+1}$. 
2408: The Lie group $G=\Gl_n(\CC)\times \Gl_1(\CC)^{n+1}$ acts on $M_{n,n+1}$ by left matrix multiplication of $\Gl_n(\CC)$ and multiplication of the $j$th factor $\Gl_1(\CC)=\CC^*$ on the $j$th column of members of $M_{n,n+1}$.
2409: By Lemma~\ref{9}, $\Der(-\log D)$ is generated by the infinitesimal action of the Lie algebra of $G$ and hence a basis of $\Der(-\log D)$ is extracted from the set of $n^2+n+1$ vector fields
2410: \begin{align}\label{22}
2411: \xi _{i,j}= \sum _{k=1}^{n+1}x_{i,k}\p_{j,k}, \text{ for } 1\leq i,j \leq n, \\
2412: \nonumber\xi _{i}= \sum _{l=1}^{n}x_{l,i}\p_{l,i}, \text{ for } 1\leq i \leq n+1,
2413: \end{align} 
2414: by omitting one because of the relation
2415: \[
2416: \chi=\sum_{i=1}^n\xi _{i,i} = \sum_{j=1}^{n+1}\xi _{j}
2417: \]
2418: corresponding to the Lie algebra of the kernel of the action.
2419: Note that the vector field $\xi _{i,i}$ resp.\ $\xi _{j}$ is the Euler vector field related to the $i$th row resp.\ to the $j$th column of the general $n\times(n+1)$-matrix and that $\chi$ is the global Euler vector field on $M_{n,n+1}$.
2420: 
2421: Since the determinant $\Delta _j$ has degree one with respect to each line and to each column except the $j$th column for which the degree is zero, the degree of $\Delta$ equals $n+1$ with respect to a row and $n$ with respect to a column. 
2422: These considerations yield
2423: \[
2424: \xi_{i,i}(\Delta)=(n+1)\Delta,\quad
2425: \xi_{j}(\Delta)=n\Delta,\quad
2426: \xi_{i,j}(\Delta)=0 \text{ for } i\ne j,
2427: \]
2428: and one can easily derive a basis of the vector fields annihilating $\Delta$.
2429: 
2430: The following lemma is self evident by definition of the action of $G$ on $M_{n,n+1}$ and we shall use it implicitly.
2431: In particular, the rank of a $G$-orbit is well-defined as the rank of any of its elements.
2432: 
2433: \begin{lemm}\
2434: \begin{asparaenum}
2435: \item Two matrices in $M_{n,n+1}$ having the same row space are in the same $G$-orbit.
2436: Similarly two matrices given by lists of column vectors $A= C_1,\dots,C_{n+1}$ and $A'=C'_1,\dots,C'_{n+1}$ are in the same $G$-orbit if there is a $\lambda _j\in\CC^*$ such that $C'_j=\lambda_jC_j$ for all $j=1,\dots,n+1$.
2437: \item If $A$ and $A'$ are in the same $G$-orbit in $M_{n,n+1}$ then any submatrix of $A$ consisting of columns $C_{i_1}, \dots , C_{i_p}$ has the same rank as the submatrix of $A'$ consisting of the corresponding columns $C'_{i_1}, \dots , C'_{i_p}$ of $A'$. 
2438: In particular $A$ and $A'$ have the same rank.
2439: \end{asparaenum}
2440: \end{lemm}
2441: 
2442: By the left action of $\Gl_n(\CC)\subseteq G$, any $G$-orbit in rank $r$ 
2443: contains, up to permutation of columns, an element of the form
2444: \beq\label{63}
2445: \begin{pmatrix}
2446: 1 & \hdots & 0 & x_{1,r+1} & \hdots & x_{1,n+1} \\
2447: \vdots & \ddots & \vdots & \vdots && \vdots \\
2448: 0 & \hdots & 1 & x_{r,r+1} & \hdots & x_{r,n+1} \\
2449: 0 & \hdots & 0 & 0 & \hdots & 0 \\
2450: \vdots && \vdots & \vdots && \vdots \\
2451: 0 & \hdots & 0 & 0 & \hdots & 0 \\
2452: \end{pmatrix}
2453: \eeq
2454: By using also the action of $G$, we may assume that $x_{i,r+1}\in\{0,1\}$:
2455: If $x_{i,r+1}\ne0$ then one can divide the $i$th row by $x_{i,r+1}$ and multiply the $i$th column by $x_{i,r+1}$.
2456: Thus there is only a finite number of maximal rank $G$-orbits including the generic orbit for which all $x_{i,n+1}$ equal $1$. 
2457: 
2458: \begin{prop}\label{64}\
2459: \begin{asparaenum}
2460: \item There are only finitely many $G$-orbits in $M_{2,3}$, and the linear free divisor $D\subseteq M_{2,3}$ is locally quasihomogeneous\footnote{See Remark \ref{26} below.}.
2461: \item The number of $G$-orbits in the linear free divisor $D\subseteq M_{3,4}$ is infinite.
2462: In particular, the set of $G$-orbits in $D$ is not locally finite, and $D$ is not Koszul free and hence not locally quasihomogeneous.
2463: \end{asparaenum}
2464: \end{prop}
2465: 
2466: \begin{proof}\
2467: \begin{asparaenum}
2468: \item The first statement follows, by Gabriel's theorem \cite{Gab72},
2469: from the fact that we are considering the representation 
2470: space of a Dynkin quiver, here of type $D_4$.
2471: In fact, in the case of $M_{2,3}$, the only orbits which 
2472: remain to be considered 
2473: are $\{0\}$ and the rank one orbits which contain, up to permutation of 
2474: columns, one of the typical elements:
2475: \[ 
2476: \begin{pmatrix}
2477: 1&1&1\\
2478: 0&0&0                       
2479: \end{pmatrix}, 
2480: \begin{pmatrix}
2481: 1&1&0\\
2482: 0&0&0                       
2483: \end{pmatrix},
2484: \begin{pmatrix}
2485: 1&0&0\\
2486: 0&0&0                        
2487: \end{pmatrix}.
2488: \] 
2489: At each point $x\neq 0$, $D$ is isomorphic to the product of the germ at $x$ of the orbit $I_x$ of $x$, and the germ at $x$ of $D':=D\cap T$, where $T$ is a smooth transversal to $I_x$ of complementary dimension.
2490: Since $T$ is logarithmically transverse to $D$ in the neighborhood of $x$, $D'$ is a free divisor.
2491: By the Cancellation Property for products of analytic spaces \cite{HM90}, $(D',x)$ is determined up to isomorphism by the fact that $(D,x)\simeq I_x\times(D',x)$, so it does not matter which transversal to $I_x$ we choose. 
2492: In the Table~\ref{90}, we take $T$ to be affine. 
2493: The local equations of $D$ shown in the last column are simply the restriction of the original equation of $D$ to the transversal
2494: $T$. 
2495: By inspection of these equations, $D$ is locally quasihomogeneous.
2496: 
2497: \begin{longtable}{|c|c|l|}
2498: \caption{Local analysis of $G$-orbits in $M_{2,3}$}\label{90}\\
2499: \hline
2500: \rule{0pt}{11pt}Representative & Transversal & Local Equation\\
2501: \hline
2502: $\begin{pmatrix}1&0&0\\0&0&1\end{pmatrix}$&
2503: $\begin{pmatrix}1&x_{12}&0\\0&x_{22}&1\end{pmatrix}$&
2504: $x_{21}x_{22}=0$\\
2505: \hline
2506: $\begin{pmatrix}1&1&1\\0&0&0\end{pmatrix}$&
2507: $\begin{pmatrix}1&1&1\\x_{21}&x_{22}&0\end{pmatrix}$&
2508: $x_{21}x_{22}(x_{22}-x_{21})=0$\\
2509: \hline
2510: $\begin{pmatrix}1&1&0\\0&0&0\end{pmatrix}$&
2511: $\begin{pmatrix}1&1&x_{13}\\0&x_{22}&x_{23}\end{pmatrix}$&
2512: $x_{22}x_{23}(x_{23}-x_{22}x_{13})=0$\\
2513: \hline
2514: $\begin{pmatrix}1&0&0\\0&0&0\end{pmatrix}$&
2515: $\begin{pmatrix}1&x_{12}&x_{13}\\0&x_{22}&x_{23}\end{pmatrix}$&
2516: $x_{22}x_{23}(x_{12}x_{23}-x_{22}x_{13})=0$\\
2517: \hline
2518: \end{longtable}
2519: 
2520: \item 
2521: In the case of $M_{3,4}$, consider the stratum in $D$ consisting of matrices of rank 2. 
2522: The four columns span a $2$-dimensional plane, and assuming they are pairwise independent, determine four lines in this plane.
2523: The cross ratio of these four lines is a $G_D$ invariant:
2524: quadruples spanning the same plane, but with different cross-ration, cannot be equivalent. 
2525: Thus there are infinitely many orbits.
2526: Now by \cite[4.3]{CN02} $D$ is not locally quasihomogeneous.
2527: \end{asparaenum}
2528: \end{proof}
2529: 
2530: \begin{prop}
2531: The linear free divisor $D\subseteq M_{n,n+1}$ from Example~\ref{41} is 
2532: strongly Euler homogeneous for any $n$.
2533: \end{prop}
2534: 
2535: \begin{proof}
2536: Let us consider a rank $r$ orbit of $G$ in $M_{n,n+1}$.
2537: If $r<n$, we can find a point $A$ in this orbit with a zero row, say row number $i$.
2538: Then the Euler vector field $\xi_{i,i}$ of this row is an Euler vector field at $A$. 
2539: 
2540: If $r=n$ we can assume that $A$ is of the form \eqref{63} with $x_{i,n+1}=1$ for $1\leq i\leq s$ and 
2541: $x_{i,n+1}=0$ for $s+1\leq i \leq n$ for some $s\leq n$.
2542: Then by \eqref{22} the space parametrized by the variables $x_{i,n+1}$ with $s+1\leq i \leq n$ is a smooth transversal to the orbit at $A$ and the restricted equation of $D$ is just $x_{s+1,n}\cdots x_{n,n+1}=0$.
2543: Thus $D$ is normal crossing and hence strongly Euler homogeneous. 
2544: \end{proof}
2545: 
2546: 
2547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2548: \subsection{Example~\ref{40} again}\label{65}
2549: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2550: 
2551: In this subsection, we show that the linear free divisors in Example \ref{40} are locally quasihomogeneous and hence Koszul free by \cite[4.3]{CN02}.
2552: By \cite{CMN96}, this implies that LCT holds although the defining group is not reductive.
2553: 
2554: We denote by $x_{i,j}$, $1\leq  i\leq j\leq n$, the coordinates on the 
2555: space of symmetric $n\times n$-matrices $\Sym_n(\CC)\subseteq M_{n,n}$. 
2556: Let $D\subseteq\Sym_n(\CC)$ be the divisor defined by the product
2557: \[
2558: \Delta={\det}_1\cdots{\det}_n
2559: \]
2560: of minors 
2561: \[
2562: {\det}_k=
2563: \begin{vmatrix}
2564: x_{1,1} & \cdots & x_{1,k} \\
2565: \vdots & & \vdots \\
2566: x_{k,1} & \cdots & x_{k,k}
2567: \end{vmatrix}.
2568: \]
2569: 
2570: By Example~\ref{40}, the group $B_n\subseteq\Gl_n(\CC)$ of upper triangular matrices acts on $\Sym_n(\CC)$ by transpose conjugation
2571: \[
2572: B\cdot S=B^tSB,\text{ for }B\in B_n, S\in\Sym_n(\CC)
2573: \]
2574: and the discriminant $D$ is a linear free divisor.
2575: Thus, $\Der(-\log D)$ can be identified with the Lie algebra of $B_n$ and has a basis consisting of the $\frac{1}{2}n(n+1)$ vector fields
2576: \[
2577: \xi_{i,j}=x_{1,i}\del 1 j +\cdots+x_{i,i}\del i j +\cdots+2x_{i,j}\del j j +
2578: \cdots+x_{i,n}\del j n \text{ for }1\leq i\le j\leq n.
2579: \]
2580: It may be helpful to view this as the symmetric matrix
2581: \[
2582: \begin{pmatrix}
2583: 0&\cdots&0&\cdots&0&x_{1,i}&0&\cdots&0\\
2584: \vdots&&\vdots&&\vdots&\vdots&\vdots&&\vdots\\
2585: 0&\cdots&0&\cdots&0&x_{i,i}&0&\cdots&0\\
2586: \vdots&&\vdots&&\vdots&\vdots&\vdots&&\vdots\\
2587: 0&\cdots&0&\cdots&0&x_{i,j-1}&0&\cdots&0\\
2588: x_{1,i}&\cdots&x_{i,i}&\cdots&x_{i,j-1}&2x_{i,j}&x_{i,j+1}&
2589: \cdots&x_{i,n}\\
2590: 0&\cdots&0&\cdots&0&x_{i,j+1}&0&\cdots&0\\
2591: \vdots&&\vdots&&\vdots&\vdots&\vdots&&\vdots\\
2592: 0&\cdots&0&\cdots&0&x_{i,n}&0&\cdots&0
2593: \end{pmatrix}
2594: \]
2595: in which all the nonzero elements lie in the $j$'th row and the $j$'th column.
2596: Note that the Euler vector field is
2597: \[
2598: \chi=\frac{1}{2}\sum_{i=1}^n\xi _{i,i}.
2599: \]
2600: For $i<j$, $\xi_{i,j}$ is nilpotent, so that $\xi_{i,j}(\Delta)=0$. 
2601: The vector field $\xi_{i,i}$ is the infinitesimal generator of the $\CC^*$ action in which the $i$'th row and column are simultaneously multiplied by $\lambda\in\CC^*$. 
2602: It follows that each determinant $\det_k$ with $k\geq i$ is homogeneous of degree $2$ with respect to $\xi_{i,i}$, and we conclude that
2603: \[
2604: \xi _{i,i}(\Delta)=2(n-i+1)\Delta, \quad \xi _{i,j}(\Delta)=0 \text{ for } i<j.
2605: \]
2606: 
2607: \begin{lemm}\label{66}
2608: There are finitely many $B_n$-orbits in $\Sym_n(\CC)$. 
2609: \end{lemm}
2610: 
2611: \begin{proof}
2612: If $i\leq j$, the pair of elementary row and column operations (``add $c$ times column $i$ to column $j$'', ``add $c$ times row $j$ to row $i$'') can be effected by the action of $B_n$. 
2613: By such operations any symmetric matrix may be brought to a normal form with at most a single nonzero element in each row and column. 
2614: Another operation in $B_n$ changes each of these nonzero elements to a $1$.
2615: Thus there are only finitely many $B_n$-orbits in $\Sym_n(\CC)$. 
2616: \end{proof}
2617: 
2618: By the discussion at the start of Section~\ref{5}, it follows that $D$ is 
2619: Koszul free. In fact this will also follow from  
2620: 
2621: \begin{prop}\label{67}
2622: The linear free divisor $D$ of Example~\ref{40} associated with the action of  $B_n$ on $\Sym _n(\CC)$ is locally quasi-homogeneous.
2623: \end{prop}
2624: 
2625: To prove this, it is enough to show that at each point $S$ of $D$ there is an element of $\Der(-\log D)_S$ which vanishes at $S$ and whose linear part is diagonal with positive eigenvalues. 
2626: This is the result of the proposition below.  
2627: In what follows we fix a symmetric matrix $S$ such that $s_{i,j}\in \{ 0,1\}$, with at most one nonzero coefficient in each row and column. 
2628: By Lemma~\ref{66}, each $B_n$ orbit contains such a matrix, and local quasihomogeneity is preserved by the $B_n$ action, so it is enough to construct a vector field of the required form in the neighborhood of each such matrix $S$.
2629: 
2630: \begin{lemm}\label{68}
2631: Assume that $s_{i,j}=1$ with $i\leq j$, then for each pair 
2632: $(k,\ell)$ in the set
2633: \[
2634: \{(i,j),(i,j+1),\dots , (i,n)\}\cup \{(i+1,j), \dots , 
2635: (j,j),(j,j+1),\dots ,(j,n)\}
2636: \]
2637: there is a vector field $v_{k,\ell}$, vanishing at $S$, 
2638: such that
2639: \begin{enumerate}[(i)]
2640: \item $v_{k,\ell}(\Delta)\in \O\cdot\Delta$, and 
2641: \item the linear part of $v_{k,\ell}$ at $S$ is equal to $(x_{k,\ell}-s_{k,\ell})\frac{\p }{\p {x_{k,\ell}}}$, and in particular is diagonal.	
2642: \end{enumerate}
2643: \end{lemm}
2644: 
2645: \begin{proof}\
2646: \begin{asparaenum}
2647: \item If $(k,\ell)=(i,\ell)$ with $j< \ell $, then
2648: \[
2649: v_{i,\ell}=x_{i,\ell}\xi_{j,\ell}= x_{i,\ell}\big[ x_{i,j}\p _{i,\ell}
2650: \mod \mm_S\Der \big]=x_{i,\ell}\p _{i,\ell}\mod 
2651: \mm^2_S\Der .
2652: \]
2653: 
2654: \item If $(k,\ell)=(i,j)$, then if $i<j$
2655: \[
2656: v_{i,j}
2657: =(x_{i,j}-1)\xi_{j,j}=(x_{i,j}-1)\big[ x_{i,j}\p _{i,j}\mod\mm_S\Der \big] 
2658: =(x_{i,j}-1)\p _{i,j}\mod \mm^2_S\Der.
2659: \]
2660: and if $i=j$
2661: \[
2662: v_{i,i}
2663: =\frac{1}{2}(x_{i,i}-1)\xi_{i,i} = (x_{i,i}-1)\p _{i,i}\mod\mm^2_S\Der.
2664: \]  
2665: 
2666: \item If $(k,\ell)=(k,j)$, with $i< k<j$ then
2667: \[
2668: v_{k,j}
2669: =x_{k,j}\xi_{i,k}= x_{k,j}\big[x_{i,j}\p _{k,j}+\mm_S\Der \big]=x_{k,j}\p _{k,j}\mod\mm^2_S\Der.
2670: \]
2671: 
2672: \item If $(k,\ell)=(j,j)$ with $i<j$, then
2673: \[
2674: v_{j,j}
2675: =\frac{1}{2}x_{j,j}\xi_{i,j}
2676: =\frac{1}{2}x_{j,j}\big[2 x_{i,j}\p _{j,j}\mod\mm_S\Der \big]
2677: =x_{j,j}\p _{j,j}\mod\mm^2_S\Der.
2678: \]
2679: 
2680: \item  If $(k,\ell)=(j,\ell)$ with $j< \ell$, then 
2681: \[
2682: v_{j,\ell}=x_{j,\ell}\xi_{i,\ell}
2683: =x_{j,\ell}\big[ x_{i,j}\p _{j,\ell}\mod\mm_S\Der \big]=x_{j,\ell}\p _{j,\ell}\mod\mm^2_S\Der.
2684: \]
2685: 
2686: \end{asparaenum}
2687: \end{proof}
2688:  
2689: \begin{lemm}\label{69}
2690: For each $i\in\{1,\dots , n\}$ there is a vector field $v_i$ vanishing at 
2691: $S$, such that 
2692: \begin{enumerate}[(i)]
2693: \item $v_i(\Delta)\in\O\cdot\Delta$
2694: \item the linear part of $v_i$ at $S$ is
2695: $
2696: \sum_{k,\ell} \lambda _{k,\ell}(x_{k,\ell}-s_{k,\ell})\frac{\p}{\p {x_{k,\ell}}}
2697: $
2698: where
2699: $\lambda _{k,\ell}=0$ if $k>i$  and 
2700: $\lambda _{i,\ell}> 0$ if $\ell \geq i$; in particular it is
2701: diagonal. 
2702: \end{enumerate}
2703: \end{lemm}
2704: 
2705: In other words we have a triangular-type system of diagonal linear 
2706: parts with positive terms on the $i$'th row 
2707: and zeros on rows after the $i$'th.
2708: 
2709: \begin{proof}
2710: If $s_{i,j}=0$ for any $j\geq i$, and $s_{k,i}=0$ for any  
2711: $k\leq i$, we can take $v_i=\xi _{i,i}$.
2712: 
2713: If $s_{k,i}=1$, with $k\leq i$, then we may apply Lemma~\ref{68} and 
2714: a linear combination of the vector fields $v_{i,i},v_{i,i+1},\dots , 
2715: v_{i,n}$ does the trick.
2716: 
2717: Finally if $s_{i,j}=1$ for some $j>i$, we observe that $\xi_{i,i}-\xi_{j,j}$, is diagonal and has non zero positive eigenvalues in the positions  
2718: \[
2719: \{(i,i),\dots,(i,j-1)\}\cup\{(i,j+1),\dots(i,n)\}
2720: .\] 
2721: Then we see that the vector field
2722: \[
2723: v_i=v_{i,j}+\xi_{i,i}-\xi_{j,j}+v_{i+1,j}+\cdots + v_{j,j}+v_{j,j+1}+\cdots+v_{j,n}
2724: \]
2725: does the trick since by adding $v_{i,j}$ we complete the row $i$ by a 
2726: positive eigenvalue at $(i,j)$, and we cancel 
2727: with the help of the appropriate $v_{k,\ell}$ all the 
2728: negative eigenvalues with row indices $k>i$.
2729: \end{proof}
2730: 
2731: \begin{prop}
2732: There is an Euler vector field $v$, $v(\Delta)\in\O\cdot\Delta$ vanishing at $S$, 
2733: with linear part diagonal and having only strictly positive eigenvalues.
2734: \end{prop}
2735: 
2736: \begin{proof}
2737: We construct $v$, by decreasing induction on $i$, as a linear combination $\alpha _nv_n+ \cdots +\alpha _1v_1$ with positive coefficients, with $\alpha _i>0$ large enough following the choice of $\alpha_n,\dots,\alpha_{i+1}$. 
2738: By construction we have $v(\Delta)=\lambda\Delta$ with $\lambda\in\O$.
2739: \end{proof}
2740: 
2741: This completes the proof of Proposition~\ref{67}.
2742: 
2743: \begin{rema}\label{26}
2744: To conclude, we mention a recent theorem of Feh\'er and Patakfalvi. 
2745: In \cite{FP07} they prove that the discriminant $D$ in the representation space of a root of a Dynkin quiver is locally quasihomogeneous. 
2746: Their theorem (\cite[Thm.~5.2]{FP07}) is stated in terms of the Incidence Property that is the subject of their paper, but their proof consists essentially of the contruction of the requisite $\CC^*$-action. 
2747: As a consequence, the LCT holds for these discriminants, by Theorem~\ref{6}.
2748: \end{rema}
2749: 
2750: 
2751: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2752: \bibliographystyle{cdraifplain}
2753: \bibliography{lfd}
2754: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2755: 
2756: \end{document}
2757: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2758: