1: \documentclass[12pt]{article}
2:
3: \textheight235mm
4: \textwidth160mm
5: \voffset-10mm
6: \hoffset-10mm
7: \parindent0cm
8: \parskip5mm
9:
10: \usepackage{amsfonts,graphicx}
11:
12:
13: \newtheorem{lemma}{Lemma}
14: \newtheorem{theorem}{Theorem}
15:
16: \title{Detecting rigid convexity of\\
17: bivariate polynomials}
18:
19: \author{Didier Henrion$^{1,2}$}
20:
21: \date{Draft of \today}
22:
23: \begin{document}
24:
25: \maketitle
26:
27: \footnotetext[1]{LAAS-CNRS, University of Toulouse, France}
28: \footnotetext[2]{Faculty of Electrical Engineering,
29: Czech Technical University in Prague, Czech Republic}
30:
31:
32: \begin{abstract}
33: Given a polynomial $x \in {\mathbb R}^n \mapsto p(x)$ in $n=2$
34: variables, a symbolic-numerical algorithm is first
35: described for detecting whether the connected component
36: of the plane sublevel set ${\mathcal P} = \{x : p(x) \geq 0\}$
37: containing the origin is rigidly convex, or equivalently,
38: whether it has a linear matrix inequality (LMI) representation,
39: or equivalently, if polynomial $p(x)$ is hyperbolic with
40: respect to the origin.
41: The problem boils down to checking
42: whether a univariate polynomial matrix is
43: positive semidefinite, an optimization problem that can be solved with eigenvalue
44: decomposition. When the variety ${\mathcal C} = \{x : p(x) = 0\}$
45: is an algebraic curve of genus zero, a second algorithm based
46: on B\'ezoutians is proposed to detect whether $\mathcal P$ has an LMI
47: representation and to build such a representation from a
48: rational parametrization of $\mathcal C$. Finally, some
49: extensions to positive
50: genus curves and to the case $n>2$ are mentioned.
51: \end{abstract}
52:
53: \begin{center}
54: {\bf\small Keywords}\\[1em]
55: Polynomial, convexity, linear matrix inequality,
56: real algebraic geometry.\\
57: \end{center}
58:
59: \section{Introduction}\label{intro}
60:
61: Linear matrix inequalities (LMIs) are versatile modeling
62: objects in the context of convex programming, with
63: many engineering applications \cite{bn}.
64: An $n$-dimensional LMI set
65: is defined as
66: \begin{equation}\label{lmi}
67: {\mathcal F} =
68: \{x \in {\mathbb R}^n \: :\: F(x) = F_0 + \sum_{i=1}^n x_i F_i
69: \succeq 0\}
70: \end{equation}
71: where the $F_i \in {\mathbb R}^{m\times m}$
72: are given symmetric matrices of size $m$
73: and $\succeq 0$ means positive semidefinite. From the
74: characteristic polynomial
75: \[
76: t \mapsto
77: \det(tI_m+F(x)) = p_0(x) + p_1(x)t + \cdots + p_{m-1}(x)t^m + t^m
78: \]
79: it follows from e.g. \cite[Theorem 20]{renegar} that
80: \begin{equation}\label{basic}
81: {\mathcal F} =
82: \{x \in {\mathbb R}^n \: :\: p_0(x) \geq 0, \ldots, p_{m-1}(x) \geq 0 \}.
83: \end{equation}
84: Hence the LMI set $\mathcal F$
85: is basic semialgebraic: it is the intersection
86: of polynomial sublevel sets. From
87: linearity of $F(x)$ and convexity of the cone of
88: positive semidefinite matrices, it also follows
89: that $\mathcal F$ is convex.
90: Hence LMI sets are {\em convex basic semialgebraic}.
91:
92:
93: \begin{figure}[h!]
94: \begin{center}
95: \includegraphics[width=15cm]{tv_screen}\\
96: \caption{The TV screen level set is not LMI.\label{tv_screen}}
97: \end{center}
98: \end{figure}
99:
100: One may then wonder whether all convex basic semialgebraic
101: sets are LMI. In \cite{hv}, Helton and Vinnikov answer
102: by the negative, showing that in the plane ($n=2$)
103: some convex basic semialgebraic sets cannot be LMI.
104: An elementary example is the so-called TV screen set
105: defined by the Fermat quartic
106: \begin{equation}\label{tv}
107: \{x \in {\mathbb R}^2 \: :\: 1-x_1^4-x_2^4 \geq 0\}
108: \end{equation}
109: see Figure \ref{tv_screen}.
110:
111: \subsection{Rigid convexity}
112:
113: Assume that the set $\mathcal F$ defined in (\ref{lmi}) has a non-empty
114: interior, and choose a point $x_0$ in this interior, i.e.
115: \[
116: x_0 \in \mathrm{int}\:{\mathcal F} = \{x \: :\:
117: F(x) \succ 0\}
118: \]
119: where $\succ 0$ means positive definite.
120: A segment starting from $x_0$ attains the boundary of $\mathcal F$
121: when the determinant $p_0(x) = \mathrm{det}\:F(x)$
122: vanishes. The remaining polynomial inequalities
123: $p_i(x) \geq 0$, $i>0$ only isolate the convex connected component
124: containing $x_0$.
125: This motivated Helton and Vinnikov \cite{hv} to study
126: semialgebraic sets defined
127: by a single polynomial inequality
128: \begin{equation}\label{poly}
129: {\mathcal P} = \{x \in {\mathbb R}^n \: :\: p(x) \geq 0\}.
130: \end{equation}
131: The set $\{x \: :\: p(x) > 0\}$ is called an algebraic interior
132: with defining polynomial $p(x)$, and it is equal to $\mathrm{int}\:{\mathcal P}$
133: when $\mathcal P$ is convex. With these notations,
134: the question addressed in \cite{hv}
135: is as follows: what are the conditions satisfied
136: by a polynomial $p(x)$ so that $\mathcal P$
137: is an LMI set ?
138:
139: For notational simplicity we will assume, without loss of generality,
140: that $x_0=0$, so that $\mathcal P$ contains the origin,
141: and hence we can normalize $p(x)$ so that $p(0)=1$.
142:
143: If $p(x) = \mathrm{det}\:F(x)$ for
144: some matrix mapping $F(x)$ we say that $p(x)$ has
145: a determinantal representation. In particular,
146: the polynomial $p_0(x)$ in (\ref{basic}) has
147: a symmetric linear determinantal representation.
148:
149: Consider an LMI set $\mathcal F$ as in (\ref{lmi}) and define
150: \[
151: p(x) = \det F(x)
152: \]
153: as the determinant of the symmetric pencil $F(x)$.
154: Note that $\deg p = m$, the dimension of $F(x)$.
155: Define the algebraic variety
156: \begin{equation}\label{var}
157: {\mathcal C} = \{x \in {\mathbb R}^n \: :\: p(x) = 0\}
158: \end{equation}
159: and notice that the boundary of $\mathcal F$ is
160: included in $\mathcal C$. Indeed, a point $x^*$
161: along the boundary of $\mathcal F$ is such that
162: the rank of $F(x^*)$ vanishes. Since the origin
163: belongs to $\mathcal F$ it holds $F_0 \succeq 0$.
164:
165: Now consider a line passing through the origin,
166: parametrized as $x(t,z) = t z$ where $t \in \mathbb R$
167: is a parameter and $z \in {\mathbb R}^n$ is any
168: vector with unit norm. For all $z$, the symmetric matrix
169: $F(x(t,z)) = F_0 + t(z_1 F_1 + \cdots + z_n F_n)$
170: has only real eigenvalues as a pencil of $t$, and its
171: determinant $t \mapsto p(x(t,z)) = \det F(x(t))$ has only
172: real roots. Therefore, a given polynomial level set
173: $\mathcal P$ as in (\ref{poly}) is LMI only if the
174: polynomial $t \mapsto p(x(t,z))$ has only real roots for all $z$,
175: it must satisfy the so-called {\em real zero condition} \cite{hv}.
176: Geometrically it means that a generic line passing through
177: the origin must intersect the variety (\ref{var})
178: at $m = \deg p$ real points. The set $\mathcal P$
179: is then called {\em rigidly convex}, a geometric property
180: implying convexity.
181:
182: A striking result of \cite{hv} is that rigid convexity
183: is also a sufficient condition for a polynomial level set
184: to be an LMI set in the plane, i.e. when $n=2$. For example,
185: it can be checked easily that the TV screen set
186: (\ref{tv}) is not rigidly convex since a generic
187: line cuts the quartic curve only twice.
188:
189: In the litterature on partial differential equations,
190: polynomials satisfying real zero condition are also
191: called hyperbolic polynomials, and the corresponding
192: LMI set is called the hyperbolicity cone, see \cite{renegar}
193: for a survey, and \cite{lpr} for connections between
194: real zero and hyperbolic polynomials.
195:
196: In passing, note the fundamental distinction between
197: an LMI set (as defined above) and a semidefinite representable
198: set, as defined in \cite{nn,bn}. A semidefinite representable
199: set is the projection of an LMI set:
200: \[
201: {\mathcal F} =
202: \{x \in {\mathbb R}^n \: :\: \exists u \in {\mathbb R}^{n_u}
203: \: :\: F(x,u) = F_0 + \sum_{i=1}^n x_i F_i + \sum_{j=1}^{n_u} u_j G_j
204: \succeq 0\}
205: \]
206: where the variables $u_j$, sometimes called liftings, are
207: instrumental to the construction of the set through
208: an extended pencil $F(x,u)$. Such a set is called a
209: lifted LMI set. It is convex semialgebraic, but in general
210: it is not basic. However, it can be expressed as a union of basic
211: semialgebraic sets. In the case of the TV screen set
212: (\ref{tv}) a lifted LMI representation follows from
213: the extended pencil
214: \[
215: F(x,u) = \left[\begin{array}{cccccc}
216: 1+u_1 & u_2 \\
217: u_2 & 1-u_1 \\
218: & & 1 & x_1 \\
219: & & x_1 & u_1 \\
220: & & & & 1 & x_2 \\
221: & & & & x_2 & u_2
222: \end{array}\right]
223: \]
224: obtained by introducing two liftings.
225: It seems that the problem of knowing which convex
226: semialgebraic sets are semidefinite representable
227: is still mostly open, see \cite{lasserre,hn}
228: for recent developments.
229:
230: \subsection{Determinantal representation}
231:
232: Once rigid convexity of a plane set, or equivalently the
233: real zero property of its defining polynomial, is established,
234: the next step is constructing an LMI representation.
235: Algebraically, given a real zero bivariate polynomial $p(x_1,x_2)$
236: of degree $m$,
237: the problem consists in finding symmetric matrices $F_0$, $F_1$
238: and $F_2$ of dimension $m$ such that
239: \[
240: p(x_1,x_2) = \det (F_0 + F_1 x_1 + F_2 x_2)
241: \]
242: and $F_0 \succeq 0$. If the $F_i$ are symmetric complex-valued
243: matrices, this is a well-studied problem
244: of algebraic geometry called {\em determinantal representation},
245: see \cite{room} for a classical reference
246: and \cite{beauville,piontkowski} for more recent surveys
247: and extensions to trivariate polynomials.
248:
249: If one relaxes the dimension constraint (allowing the $F_i$
250: to have dimension larger than $m$) and the symmetry constraint
251: (allowing the $F_i$ to be non-symmetric), then results from
252: linear systems state-space realization theory (in particular
253: linear fractional representations, LFRs) can be invoked to design
254: computer algorithms solving constructively
255: the determinantal representation problem. For example,
256: the LFR toolbox for Matlab \cite{lfr} is a user-friendly
257: package allowing to find non-symmetric determinantal
258: representations:
259: \begin{verbatim}
260: >> lfrs x1 x2
261: >> f=1/(1-x1^4-x2^4)
262: ..
263: LFR-object with 1 output(s), 1 input(s) and 0 state(s).
264: Uncertainty blocks (globally (8 x 8)):
265: Name Dims Type Real/Cplx Full/Scal Bounds
266: x1 4x4 LTI r s [-1,1]
267: x2 4x4 LTI r s [-1,1]
268: \end{verbatim}
269: The software builds a state-space realization
270: of order 8 of the transfer function $f(x)=1/p(x)$.
271: This indicates that a non-symmetric real pencil $F(x)$
272: of dimension 8 could be found that satisfies
273: $\det F(x)=p(x)$, as evidenced by the following script
274: using the Symbolic Math Toolbox
275: (Matlab gateway to Maple):
276: \begin{verbatim}
277: >> syms x1 x2
278: >> D=diag([ones(1,4)*x1 ones(1,4)*x2]);
279: >> F=eye(8)-F.a*D
280: F =
281: [ 1, -x1, 0, 0, 0, 0, 0, 0]
282: [ 0, 1, -x1, 0, 0, 0, 0, 0]
283: [ 0, 0, 1, -x1, 0, 0, 0, 0]
284: [ -x1, 0, 0, 1, -x2, 0, 0, 0]
285: [ 0, 0, 0, 0, 1, -x2, 0, 0]
286: [ 0, 0, 0, 0, 0, 1, -x2, 0]
287: [ 0, 0, 0, 0, 0, 0, 1, -x2]
288: [ -x1, 0, 0, 0, -x2, 0, 0, 1]
289: >> det(F)
290: ans =
291: -x2^4+1-x1^4
292: \end{verbatim}
293: Note that LFR and state-space realization techniques
294: are not restricted to the bivariate case, but they
295: result in pencils of large dimension (typically much larger
296: than the degree of the polynomial), and there is apparently
297: no easy way to reduce the size of a pencil.
298:
299: If one insists on having the $F_i$ symmetric, then
300: results from non-commutative state-space realizations
301: can be invoked to derive a determinantal representation,
302: at the price of relaxing the sign constraint on $F_0$.
303: An implementation is available in the {\tt NC}
304: Mathematica package \cite{hmv}. Here too, these techniques
305: may produce pencils of large dimension.
306:
307: Now if one insists on having symmetric $F_i$ of minimal
308: dimension $m$, then two essentially equivalent constructive
309: procedures are known in the bivariate case to derive Hermitian
310: complex-valued $F_i$ from a defining polynomial $p(x_1,x_2)$
311: of degree $m$. Real symmetric solutions must then be extracted
312: from the set of complex Hermitian solutions.
313:
314: The first one is based on the construction of a basis
315: for the Riemann-Roch space of complete linear systems
316: of the algebraic plane curve $\mathcal C$ given
317: in (\ref{var}). The procedure is described in \cite{dixon}:
318: one needs to find a curve of degree $m-1$ touching $\mathcal C$
319: at each intersection point, i.e. the gradients must match.
320: The algorithm is illustrated in \cite{meyerbrandis}.
321: It is not clear however how to build a touching curve
322: ensuring $F_0 \succeq 0$.
323:
324: The second determinantal representation algorithm is
325: sketched in \cite{hv} and in much more detail in \cite{vinnikov93}.
326: It is based on complex Riemann surface theory \cite{griffiths,
327: fk}. Explicit expressions for the $F_i$ matrices are given via theta
328: functions. Numerically, the key ingredient is the computation
329: of the period matrix of the algebraic curve and the
330: corresponding Abel-Jacobi map. The period matrix of a curve
331: can be computed numerically with the {\tt algcurves} package of
332: Maple, see \cite{dvh} and the tutorial
333: \cite{deconinck} for recent developments, including new
334: algorithms for explicit computations of the Abel-Jacobi map.
335: A working computer implementation taking $p(x_1,x_2)$ as input
336: and producing the $F_i$ matrices as output is still missing
337: however.
338:
339: \subsection{Contribution}
340:
341: The focus of this paper is mostly on computational
342: methods and numerical algorithms. The contribution is twofold.
343:
344: First in Section \ref{rigid} we describe an algorithm for
345: detecting rigid convexity in the plane. Given a bivariate
346: polynomial $p(x_1,x_2)$, the algorithm uses a hybrid
347: symbolic-numerical method to detect whether the
348: connected component of the sublevel set (\ref{poly})
349: containing the origin is rigidly convex.
350: The problem boils down to deciding whether a univariate
351: polynomial matrix is positive semidefinite.
352: This is a well-known problem in
353: linear systems theory, for which numerical linear
354: algebra algorithms are available (namely eigenvalue
355: decomposition), as well as a (more expensive but
356: more flexible) semidefinite programming formulation.
357:
358: Then in Section \ref{rational} we describe an algorithm
359: for solving the determinantal
360: representation problem for algebraic plane curves of genus zero.
361: The algorithm is essentially symbolic, using B\'ezoutians,
362: but it assumes that a rational parametrization of the
363: curve is available. The idea behind the algorithm is
364: not new, and can be traced back to \cite{k},
365: as surveyed recently in \cite{kaplan}. An algorithm for
366: detecting rigid convexity of a connected component delimited by
367: such curves readily follows.
368:
369: Extensions to positive genus algebraic plane curves and
370: higher dimensional sets are mentioned in Section \ref{extensions}.
371: In particular we survey the case of cubic plane curves and
372: cubic surfaces which are well understood. The case of quartic
373: (and higher degree) curves seems to be mostly open, and
374: computer implementations of determinantal representations
375: are still missing. Similarly, checking rigid convexity in
376: higher dimensions seems to be computational challenging
377: since it amounts to deciding whether a multivariate
378: polynomial matrix is positive semidefinite.
379:
380: \section{Detecting rigid convexity in the plane}\label{rigid}
381:
382: In this section we design an algorithm to assess whether the
383: connected component delimited by a bivariate polynomial around
384: the origin is rigidly convex. The idea is elementary and
385: consists in formulating algebraically the geometric condition
386: of rigid convexity of the set $\mathcal P$ defined in (\ref{poly}):
387: a line passing through the origin cuts
388: the algebraic curve $\mathcal C$ defined in (\ref{var}) a number of times which
389: is equal to the total degree $m$ of the defining bivariate polynomial
390: \[
391: x \in {\mathbb R}^2 \mapsto p(x) = \sum_{\alpha \in {\mathbb N}^2, |\alpha| \leq m}
392: p_{\alpha} x^{\alpha} =
393: p_{00} + p_{10} x_1 + p_{01} x_2 +
394: p_{20} x_1^2 + p_{11} x_1 x_2 + \cdots
395: \]
396:
397: A line passing through the origin can be
398: parametrized as:
399: \begin{equation}\label{trigo}
400: \begin{array}{rcl}
401: x_1 & = & r \cos \theta = t^{-1}(z^{-1}+z)\\
402: x_2 & = & r \sin \theta = it^{-1}(z^{-1}-z)
403: \end{array}
404: \end{equation}
405: where $z = e^{i\theta}$ and $t=2r^{-1}$.
406: Along this line, we define
407: \begin{equation}\label{pt}
408: t \in {\mathbb R} \mapsto q(t) = t^m p(x) = \sum_{k=0}^m q_k(z)t^k
409: \end{equation}
410: as a univariate polynomial of degree $m$
411: which vanishes on $\mathcal C$. Moreover $q(t)$ is monic
412: since $q_m(z) = p(0) = 1$. The remaining coefficients
413: are Laurent polynomials
414: \[
415: q_{\beta}(z) = \sum_{k=0}^m q_{\beta k}(z^k+z^{-k})
416: \]
417: with real coefficients, also called trigonometric
418: cosine polynomials.
419: Set $\mathcal P$ is rigidly convex if and only if this
420: polynomial has only real roots, i.e. if the number of
421: intersections of the line with the curve $\mathcal C$ is
422: maximal.
423:
424: \subsection{Counting the real roots of a polynomial}
425:
426: A well-known result of real algebraic geometry
427: \cite[Theorem 4.57]{bpr} states that a univariate polynomial
428: $q(t)$ of degree $m$ has only real roots if and only if
429: its Hermite matrix is positive semidefinite. The Hermite matrix
430: is the $m$-by-$m$ moment matrix of a discrete measure
431: supported with unit weights on the roots $x_1,\ldots,x_m$
432: of polynomial $q(t)$ (note that these roots are not necessarily
433: distinct). It is a symmetric Hankel matrix whose entries $(i,j)$
434: are Newton sums $N_{i+j} = \sum_{k=1}^m x_k^{i+j}$. The Newton
435: sums are elementary
436: symmetric functions of the roots that can be expressed
437: explicitly as polynomial functions of the coefficients of $q(t)$.
438: Recursive expressions are available to compute the $N_k$,
439: or equivalently, $N_k = \mathrm{trace}\:C^k$ where $C$
440: is a companion matrix of polynomial $q(t)$, i.e. a matrix
441: with eigenvalues $x_i$, see e.g. \cite[Proposition 4.54]{bpr}.
442: Recall that coefficients of the polynomial $q(t)$ given in (\ref{pt})
443: are Laurent polynomials. It follows that the Hermite matrix
444: of $q(t)$ is a symmetric trigonometric polynomial matrix of
445: dimension $m$, that we denote by $H(z)$. We have proved
446: the following result.
447:
448: \begin{lemma}\label{rh}
449: The bivariate polynomial $p(x)$ is rigidly convex
450: if and only if its Hermite matrix $H(z)$ is positive semidefinite
451: along the unit circle. Coefficients of $H(z)$
452: are explicit polynomial expressions of the coefficients of $p(x)$.
453: \end{lemma}
454:
455: \subsection{Positive semidefiniteness of polynomial matrices}
456:
457: The problem of checking positive semidefiniteness of a
458: polynomial matrix on the unit circle is generally referred
459: to as (discrete-time) spectral factorization. It is a well-known
460: problem of systems and circuit theory \cite{yakubovich,willems,glr}.
461: The positivity condition
462: can also be defined on the imaginary axis (continuous-time spectral
463: factorization) or the real axis. Various numerical methods are
464: available to solve this problem \cite{ks}. Several
465: algorithms are implemented in the Polynomial Toolbox for Matlab
466: \cite{polyx}. In increasing order of complexity, we can distinguish
467: between
468: \begin{itemize}
469: \item Newton-Raphson algorithms: the spectral factorization problem
470: is formulated as a quadratic polynomial matrix equation which is then solved
471: iteratively \cite{jk}. At each step, a linear polynomial matrix
472: equation must be solved \cite{hs}. Quadratic (resp. linear) convergence is
473: ensured locally if the polynomial matrix is positive definite (resp.
474: semidefinite);
475: \item polynomial operations: a sequence of elementary operations
476: is carried out in the ring of polynomials to reduce the
477: polynomial matrix to some canonical form, see \cite{callier}
478: and \cite{zh} for a recent survey. These algorithms are
479: cheap computationally but their numerical behavior (stability)
480: is unclear;
481: \item algebraic Riccati equation: using state-space realization,
482: the problem is formulated as a quadratic matrix equation,
483: which in turn can be solved via a matrix eigenvalue decomposition
484: with a particular structure \cite{willems,glr,tr};
485: \item semidefinite programming:
486: polynomial matrix positivity is formulated as a convex semidefinite
487: program, see \cite{tr} and the recent surveys \cite{g02,g03}.
488: The particular structure of this semidefinite program can be exploited
489: in interior-point schemes, in particular when forming
490: the gradient and Hessian. General purpose semidefinite
491: solvers can be used as well.
492: \end{itemize}
493:
494: The semidefinite programming formulation of discrete-time
495: polynomial matrix factorization, a straightfoward transposition
496: of the continuous-time case studied in \cite{tr}, is as follows.
497: The symmetric trigonometric polynomial matrix
498: $H(z)=H_0+H_1(z+z^{-1})+\cdots+H_d(z^d+z^{-d})$
499: of size $m$
500: is positive semidefinite along the unit circle
501: if and only if there is a symmetric matrix $P$
502: of size $dm$ such that
503: \begin{equation}\label{spf}
504: \begin{array}{rcl}
505: L(P) & = &
506: \left[\begin{array}{c|ccc}
507: H_0 & H_1 & \cdots & H_d \\\hline
508: H_1 & 0 & & 0 \\
509: \vdots & & \ddots \\
510: H_d & 0 & \cdots & 0
511: \end{array}\right]
512: +
513: \left[\begin{array}{ccc}
514: I \\ \hline
515: & \ddots \\
516: & & I \\
517: 0 & \cdots & 0
518: \end{array}\right]
519: P
520: \left[\begin{array}{c|ccc}
521: I & & & 0 \\
522: & \ddots & & \vdots \\
523: & & I & 0
524: \end{array}\right] \\
525: && -
526: \left[\begin{array}{ccc}
527: 0 & \cdots & 0 \\ \hline
528: I \\
529: & \ddots \\
530: & & I \\
531: \end{array}\right]
532: P
533: \left[\begin{array}{c|ccc}
534: 0 & I \\
535: \vdots & & \ddots\\
536: 0 & & & I
537: \end{array}\right] \\
538: & = &
539: \left[\begin{array}{c|c}
540: H_0 & H_{01} \\ \hline
541: H_{01}^T & 0
542: \end{array}\right]
543: +
544: \left[\begin{array}{c}
545: B^T \\ A^T
546: \end{array}\right]
547: P
548: \left[\begin{array}{c|c}
549: B & A
550: \end{array}\right]
551: -
552: \left[\begin{array}{c}
553: D^T \\ C^T
554: \end{array}\right]
555: P
556: \left[\begin{array}{c|c}
557: D & C
558: \end{array}\right]
559: \succeq 0.
560: \end{array}
561: \end{equation}
562:
563: Notice that the columns and rows of the above matrix
564: are indexed w.r.t. increasing powers of $z$
565: in such a way that
566: \[
567: B^T(z^{-1})L(P)B(z) =
568: \left[\begin{array}{c}
569: I \\ z^{-1} \\ \cdots \\ z^{-d}
570: \end{array}\right]^T
571: L(P)
572: \left[\begin{array}{c}
573: I \\ z \\ \cdots \\ z^d
574: \end{array}\right] = H(z).
575: \]
576: Positive semidefiniteness of $L(P)$ then amounts
577: to the existence of a polynomial sum-of-squares
578: decomposition of $H(z)$. From the Schur decomposition
579: $L(P) = U^T U$ with $U = \left[\begin{array}{cccc}
580: U_0 & U_1 & \cdots & U_d \end{array}\right]$
581: it follows that
582: \begin{equation}\label{pspf}
583: H(z) = U(z^{-1})^T U(z).
584: \end{equation}
585: Polynomial matrix $U(z) = U_0 + U_1 z + \cdots + U_d z^d$
586: is called a spectral factor.
587:
588: If the LMI problem (\ref{spf}) is feasible, then it
589: admits a whole family of solutions. Assuming that $H_0 \succ 0$,
590: maximizing the trace of $P$ subject to the
591: LMI constraint (\ref{spf}) yields a particular solution $P^*$
592: such that $\mathrm{rank}\:L(P^*) = m$.
593: It follows that the Schur complement of
594: \[
595: L(P) =
596: \left[\begin{array}{c|c}
597: H_0+B^TPB-D^TPD & \star \\ \hline
598: H_{01}^T+A^TPB-C^TPD & A^TPA-C^TPC
599: \end{array}\right]
600: \]
601: w.r.t. $H_0$ vanishes, where symmetric entries are denoted by $\star$.
602: This means that $P^*$ satisfies the quadratic matrix equation
603: \[
604: A^TPA-C^TPC-(H_{01}^T+A^TPB-C^TPD)H_0^{-1}(H_{01}+B^TPA-D^TPC) = 0
605: \]
606: called the (discrete-time) algebraic Riccati equation.
607: In this case, the spectral factor $U(z)$ in (\ref{pspf})
608: is square non-singular.
609:
610: \subsection{Example: cubic curve}
611:
612: Consider the component of set (\ref{poly}) around the origin
613: delimited by the cubic polynomial $p(x)=1-x_1-4x_1^2-x_2^2+4x_1^3$,
614: see Figure \ref{cubic_curve}.
615:
616: \begin{figure}[h!]
617: \begin{center}
618: \includegraphics[width=15cm]{cubic_curve}\\
619: \caption{Cubic curve and its component around the origin (shaded).\label{cubic_curve}}
620: \end{center}
621: \end{figure}
622:
623: Using the substitution (\ref{trigo}) we obtain
624: \[
625: q(t) = 12(z+z^{-1})+4(z^3+z^{-3})-(10+3(z^2+z^{-2}))t-(z+z^{-1})t^2+t^3.
626: \]
627: From the companion matrix
628: \[
629: C = \left[\begin{array}{ccc}
630: z+z^{-1} & 10+3(z^2+z^{-2}) & -12(z+z^{-1})-4(z^3+z^{-3}) \\
631: 1 & 0 & 0 \\
632: 0 & 1 & 0
633: \end{array}\right]
634: \]
635: we build (symbolically) the Hermite matrix
636: \[
637: H(z) = \left[\begin{array}{ccc}
638: 3 & \star & \star \\
639: z+z^{-1} & 22+7(z^2+z^{-2}) & \star \\
640: 22+7(z^2+z^{-2}) & 6(z+z^{-1})-2(z^3+z^{-3}) &
641: 250+124(z^2+z^{-2})+15(z^4+z^{-4})
642: \end{array}\right].
643: \]
644:
645: Solving (numerically) the LMI (\ref{spf}) with SeDuMi interfaced with
646: YALMIP yields the spectral factorization
647: (\ref{pspf}) with factor (in Matlab notation)
648: \[
649: U(z) = \left[\begin{array}{ccc}
650: -0.9021-0.7094z^2 & -0.5284z+0.2027z^3 & -11.7639-9.6359z^2-1.5201z^4\\
651: 0.1925z & 4.3449+1.6218z^2 & 0.7771z-0.5411z^3 \\
652: 1.1578-0.5527z^2 & 0.3819z+0.1579z^3 & 2.4331-2.8689z^2-1.1844z^4
653: \end{array}\right]
654: \]
655: which certifies numerically that $p(x)$ is rigidly convex.
656:
657: \subsection{Example: quartic curve}
658:
659: Let us apply the algorithm to test rigid convexity
660: of the TV quartic level set (\ref{tv}) with
661: $p(x)=1-x_1^4-x_2^4$, see Figure \ref{tv_screen}.
662:
663: We obtain $q(t) = -12-2(z^4+z^{-4})+t^4$ and
664: the Hermite matrix
665: \[
666: H(z) = \left[\begin{array}{cccc}
667: 4 & \star & \star & \star \\
668: 0 & 0 & \star & \star \\
669: 0 & 0 & 48+8(z^4+z^{-4}) & \star \\
670: 0 & 48+8(z^4+z^{-4}) & 0 & 0
671: \end{array}\right].
672: \]
673: From the zero diagonal entries and
674: the non-zero entries in the corresponding rows and columns
675: we conclude that $H(z)$ cannot be positive semidefinite
676: and hence that the TV quartic level set is not rigidly
677: convex.
678:
679: \subsection{Numerical considerations}
680:
681: Since the Hermite matrix $H(z)$ has a Hankel structure, and
682: positive definite symmetric Hankel matrices have a
683: conditioning (ratio of extreme eigenvalues) which can
684: be bounded below by an exponential function of the
685: matrix size \cite{becker,tyrty}, it may be appropriate
686: to apply a congruence transformation on matrix $H(z)$, also
687: called scaling.
688:
689: For example, if $H(e^{i\theta_0})$ is positive definite
690: for some $\theta_0$ (say $\theta_0=0$, but other choices
691: are also possible), it admits
692: a Schur factorization $H(e^{i\theta_0})=V^T D V$ with $V$
693: orthogonal and $D$ diagonal non-singular. If $D$ is reasonably
694: well-conditioned, we can test positive semidefiniteness of
695: the modified trigonometric polynomial matrix
696: $H_0(z)=VD^{-1/2}H(z)D^{-1/2}V^T$ along the unit circle,
697: which is such that $H_0(e^{i\theta_0})$ is the identity matrix.
698: If $D$ is not well-conditioned, we can still use
699: $H_0(z) = VH(z)V^T$ which is such that $H_0(e^{i\theta_0})$ is
700: a diagonal matrix.
701:
702: The impact of this data scaling
703: on the numerical behavior of the semidefinite programming or algebraic Riccati
704: equation solvers is however out of the scope of this paper.
705:
706: \section{LMI sets and rational algebraic plane curves}\label{rational}
707:
708: In the case that the algebraic curve $\mathcal C$ in (\ref{var})
709: has genus zero, i.e. the curve is rationally parametrizable,
710: an alternative algorithm can be devised to test rigid convexity
711: of a connected component delimited by $\mathcal C$. The algorithm
712: is based on elimination theory. It uses a particular symmetric
713: form of a resultant called the B\'ezoutian. As a by-product, the
714: algorithm also solves the determinantal representation problem
715: in this case. As surveyed recently in \cite{kaplan}, the key idea of
716: using B\'ezoutians in the context of determinantal
717: representations can be traced back to \cite{k}.
718:
719: Starting from the implicit representation
720: \begin{equation}\label{implicit}
721: {\mathcal C} = \{x \in {\mathbb R}^2 \: :\: p(x) = 0\}
722: \end{equation}
723: of curve $\mathcal C$, with $p(x)$ a bivariate polynomial
724: of degree $m$, we apply a parametrization algorithm to obtain
725: an explicit representation
726: \begin{equation}\label{param}
727: {\mathcal C} = \{x \in {\mathbb R}^2 \: :\: x_1 = q_1(u)/q_0(u), \:
728: x_2 = q_2(u)/q_0(u), \: u \in {\mathbb R}\}
729: \end{equation}
730: with $q_i(u)$ univariate polynomials of degree $m$.
731: Algorithms for parametrizing an implicit algebraic curve are
732: described in \cite{abhyankar,sendra,vanhoeij}. An implementation
733: by Mark van Hoeij
734: is available in the {\tt algcurves} package of Maple. The
735: coefficients of $q_i(u)$ are generally found
736: in an algebraic extension of small degree over the field
737: of coefficients of $p(x)$.
738:
739: With the help of resultants, we can eliminate the variable
740: $u$ in parametrization (\ref{param}) and recover an implicit
741: equation (\ref{implicit}), see \cite[Section 3.3]{cox}.
742: To address this implicitization problem,
743: we make use of a particular resultant, the B\'ezoutian, see
744: \cite[Section 5.1.2]{mourrain}.
745: Given two univariate polynomials $g, h$ of the same degree $m$
746: (if the degree is not the same, the smallest degree polynomial is considered
747: as a degree $m$ polynomial with zero leading coefficients)
748: build the following bivariate polynomial
749: \[
750: \frac{g(u)h(v)-g(v)h(u)}{u-v} = \sum_{k=0}^{m-1} \sum_{l=0}^{m-1}
751: b_{kl} u^k v^l
752: \]
753: called the B\'ezoutian of $g$ and $h$,
754: and the corresponding symmetric matrix $B(g,h)$
755: of size $m\times m$ with entries $b_{kl}$ bilinear in coefficients
756: of $g$ and $h$. As shown e.g. in
757: \cite[Section 5.1.2]{mourrain},
758: the determinant of the B\'ezoutian matrix is the resultant,
759: so we can use it to derive the
760: implicit equation (\ref{implicit}) of a curve
761: from the explicit equations (\ref{param}).
762:
763: \begin{lemma}\label{pencil}
764: Given polynomials $q_0, q_1, q_2$ in (\ref{param}), a polynomial
765: $p$ in (\ref{implicit}) is given by
766: $p(x) = \det F(x)$ where
767: \begin{equation}\label{px}
768: \begin{array}{rcl}
769: F(x) & = & B(q_1,q_2)+x_1B(q_2,q_0)+x_2B(q_1,q_0) \\
770: & = & F_0+x_1F_1+x_2F_2
771: \end{array}
772: \end{equation}
773: is a symmetric pencil of size $m$.
774: \end{lemma}
775:
776: {\bf Proof:} Rewrite the system of equations (\ref{param}) as
777: \[
778: \begin{array}{rcccl}
779: g_1(u) & = & q_1(u) - x_1 q_0(u) & = & 0 \\
780: g_2(u) & = & q_2(u) - x_2 q_0(u) & = & 0 \\
781: \end{array}
782: \]
783: and use the B\'ezoutian resultant
784: to eliminate indeterminate $u$ and obtain conditions
785: for a point $(x_1,x_2)$ to belong to the curve. The B\'ezoutian
786: matrix is $B(g_1,g_2) = B(q_1-x_1q_0,q_2-x_2q_0) =
787: B(q_1,q_2)+x_1B(q_2,q_0)+x_2B(q_1,q_0)$. Linearity in $x$
788: follows from bilinearity of the B\'ezoutian
789: and the common factor $q_0(u)$.$\Box$
790:
791: Lemma \ref{pencil} provides an implicit equation of
792: curve (\ref{implicit}) in symmetric linear determinantal form.
793:
794: \subsection{Detecting rigid convexity}
795:
796: Once polynomial $p(x)$ is in symmetric linear determinantal
797: form as in Lemma \ref{pencil}, checking rigid convexity
798: of the connected component containing the origin $x=0$
799: amounts to testing positive definiteness of $F(0) = F_0 = B(q_1,q_2)$.
800:
801: \begin{lemma}\label{definite}
802: The B\'ezoutian matrix $B(q_1,q_2)$ is positive semidefinite
803: if and only if polynomial $q_1(u)$ and $q_2(u)$
804: have only real roots that interlace.
805: \end{lemma}
806:
807: {\bf Proof:} The signature (number of
808: positive eigenvalues minus number of negative eigenvalues)
809: of the B\'ezoutian of $q_1(u)$ and $q_2(u)$
810: is the Cauchy index of the rational function $q_1(u)/q_2(u)$,
811: the number of jumps of the function from $-\infty$ to $+\infty$
812: minus the number of jumps from $+\infty$ to $-\infty$, see
813: \cite[Definition 2.53]{bpr} or \cite[Theorem 9.4]{bpr}.
814: It is maximum when $B(q_1,q_2)$ is positive definite.
815: This occurs if and only if the roots
816: of $q_1(u)$ and $q_2(u)$ are all real and interlace.$\Box$
817:
818: \begin{lemma}\label{rigidbez}
819: The connected component around the origin delimited
820: by curve (\ref{implicit}) is rigidly convex if and only if
821: $B(q_1,q_2) \succeq 0$.
822: \end{lemma}
823:
824: {\bf Proof:} Since $F_0 \succeq 0$, the set admits the LMI
825: representation $\{x \in {\mathbb R}^2 \: :\: F(x) \succeq 0\}$,
826: which is equivalent to being rigidly convex. $\Box$
827:
828: \subsection{Finding a rigidly convex component}
829:
830: If the connected component around the origin is not
831: ridigly convex, it may happen that there is another
832: rigidly convex connected component elsewhere.
833: To find it, it suffices to determine a point $x^*$
834: such that $F(x^*) \succeq 0$.
835: This is equivalent to solving
836: a bivariate LMI problem.
837:
838: We can apply primal-dual interior-point methods \cite{nn}
839: to solve this semidefinite programming problem, since
840: the function $f(x) = - \log p(x) = \log \det F(x)^{-1}$ is a strictly
841: convex self-concordant barrier for the interior of the LMI set. If the
842: LMI set is bounded, minimizing $f(x)$ yields the analytic
843: center of the set. If the LMI set
844: is empty, the dual semidefinite problem yields a
845: Farkas certificate of infeasibility. However, in the
846: bivariate case a point $x^*$ satisfying
847: $F(x^*) \succeq 0$ can be found more easily
848: with real algebraic geometry and univariate
849: polynomial root extraction.
850:
851: A first approach consists in identifying the local
852: minimizers of function $f(x)$. They are
853: such that the gradient $g(x)$ of $p(x)$ vanishes, i.e.
854: they are such that
855: \begin{equation}\label{grad}
856: g_i(x) = \frac{\partial p(x)}{\partial x_i} =
857: \mathrm{trace}(p(x)F^{-1}(x)F_i) = 0.
858: \end{equation}
859: We can characterize these minimizers by eliminating
860: one variable, say $x_1$, from the system $g_1(x)=g_2(x)=0$,
861: and solving for the other variable $x_2$ via
862: polynomial root extraction. Resultants
863: can be used for that purpose.
864:
865: A second approach consists in finding points on
866: the boundary of the LMI set, which are such that
867: $p(x)=0$ and either $g_1(x)=0$ or $g_2(x)=0$.
868: Here too, resultants can be applied to end up
869: with a polynomial root extraction problem.
870:
871: From the points generated by these two procedures,
872: we keep only those satisfying $F(x) \succeq 0$,
873: an inequality that can be certified by testing
874: the signs of the coefficients of the characteristic
875: polynomial of $F(x)$, as explained in the
876: introduction.
877:
878: \subsection{Example: capricorn curve}
879:
880: Let $p(x) = x_1^2(x_1^2+x_2^2)-2(x_1^2+x_2^2-x_2)^2$.
881: With the {\tt parametrization} function of the {\tt algcurves}
882: package of Maple, we obtain a rational parametrization
883: \[
884: \begin{array}{rcl}
885: q_0(t) & = & 45-8t+10t^2+t^4 \\
886: q_1(t) & = & -7+44t-18t^2-4t^3+t^4\\
887: q_2(t) & = & 49-28t-10t^2+4t^3+t^4.
888: \end{array}
889: \]
890: With the {\tt BezoutMatrix} function of the {\tt LinearAlgebra}
891: package, we build the corresponding symmetric pencil
892: \[\small
893: F(x) = \left[\begin{array}{cccc}
894: 8-4x_1-4x_2 & \star & \star & \star \\
895: 8+20x_1-28x_2 & 40+60x_1+92x_2 & \star & \star \\
896: -72+20x_1+52x_2 & -8-36x_1-84x_2 & 776+540x_1+476x_2 & \star \\
897: 56-4x_1-52x_2 & -168+180x_1+180x_2 & -952-940x_1+740x_2 & 1960-868x_1-1924x_2
898: \end{array}\right]
899: \]
900: whose determinant (up to a constant factor) is equal to $p(x)$.
901: The eigenvalues of $F(0)$ are equal to $0$ (double)
902: and $1392\pm48\sqrt{533}$. They are all non-negative
903: which indicates that the origin lies on the boundary of
904: an LMI region defined by $F(x) \succeq 0$.
905:
906: Values of $x_2$ at local optima satisfying the system
907: of cubic equations (\ref{grad})
908: can be found with Maple as follows:
909: \begin{verbatim}
910: > p:=x1^2*(x1^2+x2^2)-2*(x1^2+x2^2-x2)^2:
911: > solve(resultant(diff(p,x1),diff(p,x2),x1));
912: 0, 0, 0, 1, 1/2, 3+sqrt(5), 3-sqrt(5), 3+sqrt(5), 3-sqrt(5)
913: \end{verbatim}
914: from which it follows that, say, the point $x_1=0$, $x_2=1/2$
915: is such that $F(x) \succ 0$.
916:
917: \begin{figure}[h!]
918: \begin{center}
919: \includegraphics{capricorn_curve}\\
920: \caption{Capricorn curve defining an LMI region (shaded).\label{capricorn_curve}}
921: \end{center}
922: \end{figure}
923:
924: The corresponding LMI region together with the quartic
925: capricorn curve are represented on Figure
926: \ref{capricorn_curve}.
927:
928: \subsection{Example: bean curve}
929:
930: \begin{figure}[h!]
931: \begin{center}
932: \includegraphics[width=15cm]{bean_curve}\\
933: \caption{Bean curve defining a region which is not LMI (shaded).
934: \label{bean_curve}}
935: \end{center}
936: \end{figure}
937:
938: Let $p(x)=x_1^4+x_1^2x_2^2+x_2^4-x_1^3+x_1x_2^2$.
939: With the B\'ezoutians we obtain the following pencil
940: \[
941: F(x) = \left[\begin{array}{cccc}
942: x_1& \star & \star & \star\\
943: x_2&1& \star & \star \\
944: x_1&x_2&0& \star\\
945: x_2&1-x_1&0&1-x_1
946: \end {array}
947: \right].
948: \]
949: We can check that $F(0)$ has eigenvalues $2$ and $0$ (triple)
950: and this is the only point for which $F(x) \succeq 0$.
951: It follows that the convex set delimited by the curve
952: $p(x)=0$ is not LMI, see Figure \ref{bean_curve}.
953:
954: \section{Extensions}\label{extensions}
955:
956: In this paragraph we outline some potential extensions
957: of the results to algebraic plane curves of positive
958: genus and varieties of higher dimensions.
959:
960: \subsection{Cubic plane curves}
961:
962: The case of cubic plane algebraic curves is well understood,
963: see e.g. \cite{vinnikov86} or \cite{taussky}. Singular cubics
964: (genus zero) can be handled
965: via B\'ezoutians as in Section \ref{rational}.
966: Smooth cubics (genus one), also called elliptic curves,
967: can be handled via their Hessians.
968:
969: Let $p(x_1,x_2)$ be a cubic polynomial that
970: we homogeneize to $p(x_0,x_1,x_2) = x_0^3p(x_1/x_0,x_2/x_0)$.
971: Define its 3-by-3 symmetric
972: Hessian matrix $H(p(x))$ with entries
973: \[
974: H_{ij} = \frac{\partial^2 p(x)}{\partial x_i \partial x_j}
975: \]
976: and the corresponding Hessian $h(x)=\det H(p(x))$.
977: The elliptic curve $p(x)=0$ has 9 inflection points,
978: or flexes, satisfying $p(x)=h(x)=0$, and 3 of them are real.
979: Since $p(x)$ and $h(x)$ share the same flexes and the Hessian
980: matrix yields a symmetric linear determinantal representation for $h(x)$,
981: we can use homotopy to find a determinantal representation
982: for $p(x)$.
983:
984: For real $t$ define the parametrized Hessian
985: $g(x,t)=\det H(h(x)+tp(x))$ and find $t^*$ satisfying
986: $g(x^*,t^*)=p(x^*)$ at a real flex $x^*$ by solving
987: a cubic equation. As a result, we obtain three
988: distinct symmetric pencils not equivalent by
989: congruence transformation. One of the may be definite
990: hence LMI.
991:
992: For example, let $p(x)=x_1^3-x_2^2-x_1$. Build the
993: Hessian $h(x)=\det H(p(x)) = 8(x_0^3+3x_0x_1^2-3x_1x_2^2)$
994: and the parametrized Hessian $g(x,t)=\det H(h(x)+tp(x))=
995: 24t^3x_0x_1^2-576t^2x^2_0x_1+\cdots+110592x_1^3$.
996: Polynomial $g(x,t)$ matches $g(x)$ at flex $x^*_0=0$ for
997: $t^* \in \{0, 24, -24\}$
998: yielding the following three representations
999: \[
1000: \begin{array}{rcl}
1001: F^1(x) & = & \left[\begin{array}{ccc}
1002: 1 & \star & \star \\ -x_2 & -x_1 & \star \\ x_1 & 0 & 1
1003: \end{array}\right] \\ \\
1004: F^2(x) & = & 4^{-\frac{1}{3}}\left[\begin{array}{ccc}
1005: 1+3x_1 & \star & \star \\
1006: -x_2 & -1-x_1 & \star\\
1007: -1+x_1 & -x_2 & 1-x_1
1008: \end{array}\right] \\ \\
1009: F^3(x) & = & 4^{-\frac{1}{3}}\left[\begin{array}{ccc}
1010: 1-3x_1 & \star & \star\\
1011: -x_2 & 1-x_1 & \star \\
1012: 1+x_1 & x_2 & 1+x_1
1013: \end{array}\right]
1014: \end{array}
1015: \]
1016: such that $\det F^i(x)=p(x)$ for all $i=1,2,3$.
1017: Only the first one generates an LMI set $F^1(x)\succeq 0$.
1018:
1019: \subsection{Positive genus plane curves}
1020:
1021: The case of algebraic plane curves
1022: of positive genus and degree equal to four (quartic)
1023: or higher is mostly open. Whereas rigid convexity of higher
1024: degree polynomials can be checked with the proposed approach,
1025: there is no known implementation of an algorithm that
1026: produces symmetric linear determinantal (and hence LMI)
1027: representations in this case. For quartics, contact
1028: curves can be recovered from bitangents. In \cite{edge}
1029: complex symmetric linear determinantal representations
1030: of the quartic $1+x^4_1+x^4_2$ could be derived from the
1031: equations of the bitangents found previously
1032: by Cayley for this particular curve.
1033:
1034: B\'ezoutians can be generalized to the multivariate
1035: case, as surveyed in \cite{mourrain}. In Lemma \ref{pencil}
1036: we derived a symmetric linear determinantal representation
1037: by eliminating the variable $u$ in the system of
1038: equations
1039: \[
1040: \begin{array}{rclcl}
1041: g_1(u) & = & q_1(u) - x_1 q_0(u) & = & 0 \\
1042: g_2(u) & = & q_2(u) - x_2 q_0(u) & = & 0 \\
1043: \end{array}
1044: \]
1045: corresponding to a rational parametrization
1046: $x_1(u) = q_1(u)/q_0(u)$, $x_2(u) = q_2(u)/q_0(u)$
1047: of the curve $p(x_1,x_2)=0$. In the positive genus case, such
1048: a rational parametrization is not available, but
1049: we can still define a system of equations
1050: \[
1051: \begin{array}{rclcl}
1052: g_1(u_1,u_2) & = & x_1-u_1 & = & 0 \\
1053: g_2(u_1,u_2) & = & x_2-u_2 & = & 0 \\
1054: g_3(u_1,u_2) & = & p(u_1,u_2) & = & 0
1055: \end{array}
1056: \]
1057: describing the curve $p(x_1,x_2)=0$ after eliminating
1058: variables $u_1$ and $u_2$. Define the discrete differentials
1059: \[
1060: \partial_1 g(u,v) = \frac{g(u_1,u_2)-g(v_1,v_2)}{u_1-v_1}, \quad
1061: \partial_2 g(u,v) = \frac{g(v_1,u_2)-g(v_1,v_2)}{u_2-v_2}
1062: \]
1063: and the quadratic form
1064: \[
1065: \det \left[\begin{array}{ccc}
1066: g_1 & \partial_1 g_1 & \partial_2 g_1 \\
1067: g_2 & \partial_1 g_2 & \partial_2 g_2 \\
1068: g_3 & \partial_1 g_3 & \partial_2 g_3
1069: \end{array}\right] =
1070: \det \left[\begin{array}{ccc}
1071: x_1-u_1 & -1 & 0 \\
1072: x_2-u_2 & 0 & -1 \\
1073: p(u_1,u_2) & \partial_1 p(u,v) & \partial_2 p(u,v)
1074: \end{array}\right] =
1075: \sum_{\alpha,\beta} f_{\alpha,\beta} u^{\alpha}v^{\beta}
1076: \]
1077: using bi-indices $\alpha$ and $\beta$. Then the matrix
1078: $F(x)$ of the quadratic form is a symmetric pencil
1079: satisfying $\det F(x) = p(x)q(x)$ where $q(x)$
1080: is an extraneous factor. We hope that $q(x)$ does
1081: not depend on $x$, even though this cannot be
1082: guaranteed in general. For example, in the case of the
1083: Fermat curve $p(x) = 1-x^4_1-x^4_2$ whose genus is three,
1084: using the {\tt multires} package for Maple \cite{multires},
1085: we could obtain
1086: \[
1087: F(x) = \left[\begin{array}{ccccccc}
1088: -1 & 0 & 0 & 0 & 0 & x_1 & x_2 \\
1089: 0 & 0 & 0 & x_1 & 0 & -1 & 0 \\
1090: 0 & 0 & 0 & 0 & x_2 & 0 & -1 \\
1091: 0 & x_1 & 0 & -1 & 0 & 0 & 0 \\
1092: 0 & 0 & x_2 & 0 & -1 & 0 & 0 \\
1093: x_1 & -1 & 0 & 0 & 0 & 0 & 0 \\
1094: x_2 & 0 & -1 & 0 & 0 & 0 & 0
1095: \end{array}\right]
1096: \]
1097: which is such that $\det F(x) = -p(x)$, i.e. $q(x)=-1$.
1098:
1099: \subsection{Surfaces and hypersurfaces}
1100:
1101: The case $n=m=3$, i.e. cubic surfaces, is well
1102: understood, see \cite{kosir} for a full constructive
1103: development. All the self-adjoint linear determinantal
1104: representations can be obtained from the tritangent
1105: planes. The number of non-equivalent
1106: representations depends on the number
1107: and class of real lines among the 27 complex lines
1108: of the surface. See \cite{szilagyi} for
1109: a nice survey on cubic surfaces.
1110:
1111: \begin{figure}[h!]
1112: \begin{center}
1113: \includegraphics[width=15cm]{cayley_cubic}\\
1114: \caption{The Cayley cubic surface with its convex connected
1115: component.\label{cayley_cubic}}
1116: \end{center}
1117: \end{figure}
1118:
1119: A well-known example is the Cayley cubic
1120: \[
1121: \frac{1}{u_0}+\frac{1}{u_1}+\frac{1}{u_2}+\frac{1}{u_3} = 0
1122: \]
1123: whose algebraic equation is
1124: \[
1125: u_0 u_1 u_2 + u_0 u_1 u_3 + u_0 u_2 u_3 + u_1 u_2 u_3 = 0.
1126: \]
1127: Under involutary linear mapping
1128: \[
1129: \left[\begin{array}{c}
1130: x_0 \\ x_1 \\ x_2 \\ x_3
1131: \end{array}\right] =
1132: \frac{1}{2}
1133: \left[\begin{array}{cccc}
1134: 1 & 1 & 1 & 1 \\ 1 & 1 & -1 & -1 \\ 1 & -1 & 1 & -1 \\ 1 & -1 & -1 & 1
1135: \end{array}\right]
1136: \left[\begin{array}{c}
1137: u_0 \\ u_1 \\ u_2 \\ u_3
1138: \end{array}\right]
1139: \]
1140: the dehomogenized ($x_0=1$)
1141: algebraic equation becomes
1142: \[
1143: p(x) =
1144: 1-x_1^2-x_2^2-x_3^2-2x_1x_2x_3 =
1145: \det \left[\begin{array}{ccc}1&x_1&x_2\\
1146: x_1&1&x_3\\x_2&x_3&1\end{array}\right] =
1147: \det F(x)
1148: \]
1149: which is the determinant of the 3x3 moment
1150: matrix of the MAXCUT LMI relaxation.
1151: The surface $p(x)=0$ is represented on Figure \ref{cayley_cubic},
1152: using the {\tt surf} visualization package.
1153: In particular, we can easily identify the convex connected
1154: component containing the origin, described by
1155: the LMI $F(x) \succeq 0$. The component has four vertices,
1156: or singularities, for which the rank of $F(x)$ drops down to one.
1157:
1158: In general, only curves and cubic surfaces admit
1159: generically a determinantal representation.
1160: When $n>3$ or $m>3$ and no lifting is allowed,
1161: the hypersurface $p(x)=0$ must be highly
1162: singular to have a determinantal representation
1163: \cite{beauville}, and hence, a fortiori, an LMI
1164: representation.
1165: This leaves however open the existence of
1166: alternative algorithms consisting in constructing
1167: symmetric linear determinantal representations
1168: of modified polynomials $p(x)q(x)$, with
1169: $q(x)$ globally nonnegative, say $q(x)=(\sum_i x^{2k}_i)$
1170: or $(\sum_i x_i)^{2k}$
1171: for $k\geq 1$ large enough.
1172:
1173: Finally, let us conclude by remarking that, as a
1174: by-product of the proof leading to Lemma \ref{rh},
1175: checking numerically rigid convexity of
1176: a scalar polynomial when $n>2$
1177: amounts to checking positivity of a multivariate
1178: Hermite matrix. See e.g. \cite{dumitrescu} for recent
1179: developments on the use of semidefinite programming
1180: for multivariate trigonometric polynomial matrix
1181: positivity.
1182:
1183: \section*{Acknowledgments}
1184:
1185: This work benefited from feedback by (in alphabetical order)
1186: Daniele Faenzi,
1187: Leonid Gurvits, Fr\'ed\'eric Han,
1188: Bill Helton, Toma\v z Ko\v sir, Salma Kuhlmann,
1189: Jean-Bernard Lasserre, Adrian Lewis,
1190: Bernard Mourrain, Pablo Parrilo,
1191: Jens Piontkowski, Mihai Putinar, Bernd Sturmfels,
1192: Jean Vall\`es and Victor Vinnikov.
1193:
1194: \begin{thebibliography}{XX}
1195:
1196: \bibitem{abhyankar}
1197: S. S. Abhyankar, C. L. Bajaj.
1198: Automatic parameterization of rational curves and surfaces III:
1199: Algebraic plane curves. Computer Aided Geom. Design 5:309--321,
1200: 1988.
1201:
1202: \bibitem{bpr}
1203: S. Basu, R. Pollack, M.-F. Roy. Algorithms in
1204: real algebraic geometry. Springer, 2003.
1205:
1206: \bibitem{beauville}
1207: A. Beauville.
1208: Determinantal hypersurfaces.
1209: Michigan Math. J. 8:39--64, 2000.
1210:
1211: \bibitem{becker}
1212: B. Beckermann.
1213: The condition number of real Vandermonde, Krylov and positive definite
1214: Hankel matrices. Numer. Mathematik, 85:553-577, 2000.
1215:
1216: \bibitem{bn}
1217: A. Ben-Tal, A. Nemirovskii. Lectures on modern convex optimization.
1218: SIAM, 2001.
1219:
1220: \bibitem{multires}
1221: L. Bus\'e, B. Mourrain, I.Z. Emiris, O. Ruatta, J. Canny, P. Pedersen,
1222: I. Tonelli. Using the Maple {\tt multires} package.
1223: INRIA Sophia-Antipolis, France, 2003.
1224:
1225: \bibitem{kosir}
1226: A. Buckley, T. Ko\v sir.
1227: Determinantal representations of smooth cubic surfaces.
1228: Geometriae Dedicata, 125:115--140, 2007.
1229:
1230: \bibitem{callier}
1231: F. M. Callier. On polynomial matrix spectral factorization
1232: by symmetric extraction. IEEE Trans. Autom. Control,
1233: 30(5):453--464, 1985.
1234:
1235: \bibitem{cox}
1236: D. Cox, J. Little, D. O'Shea. Ideals, varietes, and
1237: algorithms. Springer, 1992.
1238:
1239: \bibitem{deconinck}
1240: B. Deconinck, M. S. Patterson.
1241: Computing the Abel map. Preprint, Dept. Applied Math.,
1242: Univ. Washington, Seattle, 2007.
1243:
1244: \bibitem{dvh}
1245: B. Deconinck, M. van Hoeij.
1246: Computing Riemann matrices of algebraic curves. Phys. D
1247: 152/153:123--152, 2001.
1248:
1249: \bibitem{dixon}
1250: A. C. Dixon. Note on the reduction of a ternary quantic to a
1251: symmetrical determinant. Proc. Cambridge Phil.
1252: Soc. 11:350--351, 1902.
1253:
1254: \bibitem{dumitrescu}
1255: B. Dumitrescu. Positive trigonometric polynomials and signal
1256: processing applications. Springer, 2007.
1257:
1258: \bibitem{edge}
1259: W. L. Edge. Determinantal representations of $x^4+y^4+z^4$.
1260: Proc. Cambridge Phil. Soc. 34:6--21, 1938.
1261:
1262: \bibitem{mourrain}
1263: M. Elkadi, B. Mourrain.
1264: Introduction \`a la r\'esolution des syst\`emes polynomiaux.
1265: Springer, 2007.
1266:
1267: \bibitem{fk}
1268: H. M. Farkas, I. Kra. Riemann surfaces. 2nd ed. Springer, 1992.
1269:
1270: \bibitem{g02}
1271: Y. Genin, Y. Hachez, Y. Nesterov, R. \c{S}tefan, P. Van Dooren, S. Xu.
1272: Positivity and linear matrix inequalities. European J.
1273: Control, 8(3):275--298, 2002.
1274:
1275: \bibitem{g03}
1276: Y. Genin, Y. Hachez, Y. Nesterov, P. Van Dooren.
1277: Optimization problems over positive pseudo-polynomial matrices.
1278: SIAM J. Matrix Anal. Appl., 25(1):57--79, 2003.
1279:
1280: \bibitem{glr}
1281: I. Gohberg, P. Lancaster, I. Rodman.
1282: Matrix polynomials, Academic Press, 1982.
1283:
1284: \bibitem{griffiths}
1285: P. A. Griffiths. Introduction to algebraic curves.
1286: AMS, 1989.
1287:
1288: \bibitem{lfr}
1289: S. Hecker, A. Varga, J. F. Magni.
1290: Enhanced LFR toolbox for Matlab.
1291: Proc. IEEE Symp. Computer Aided Control System Design,
1292: Taiwan, 2004.
1293:
1294: \bibitem{hmv}
1295: J. W. Helton, S. A. McCullough, V. Vinnikov.
1296: Noncommutative convexity arises from linear matrix inequalities.
1297: J. Functional Analysis 240(1):105--191, 2006.
1298:
1299: \bibitem{hv}
1300: J. W. Helton, V. Vinnikov. Linear matrix inequality representation of
1301: sets. Comm. Pure Applied Math. 60(5):654--674, 2007.
1302:
1303: \bibitem{hn}
1304: J. W. Helton, J. Nie.
1305: Sufficient and necessary conditions for semidefinite representability
1306: of convex hulls and sets.
1307: {\tt arXiv:0709.4017}, September 2007.
1308:
1309: \bibitem{hs}
1310: D. Henrion, M. \v Sebek.
1311: An efficient numerical method for the discrete-time symmetric matrix
1312: polynomial equation. IEE Proc. Control Theory Appl. 145(5):443--448, 1998.
1313:
1314: \bibitem{jk}
1315: J. Je\v zek, V. Ku\v cera. Efficient algorithm for matrix spectral
1316: factorization. Automatica, 21(6):663-669, 1985.
1317:
1318: \bibitem{kaplan}
1319: S. Kaplan, A. Shapiro, M. Teicher.
1320: Several applications of B\'ezout matrices.
1321: {\tt arXiv:math.AG/0601047}, January 2006.
1322:
1323: \bibitem{k}
1324: N. Kravistky. On the discriminant function of two commuting non-selfadjoint
1325: operators. Integral Equ. Operator Theory, 3(1):97--124, 1980.
1326:
1327: \bibitem{ks}
1328: H. Kwakernaak, M. \v Sebek. Polynomial J-Spectral Factorization.
1329: IEEE Trans. Autom. Control, 39(2):315--328, 1994.
1330:
1331: \bibitem{lasserre}
1332: J. B. Lasserre. Convex sets with lifted semidefinite representation.
1333: LAAS-CNRS Research Report No. 07034, January 2007.
1334:
1335: \bibitem{lpr}
1336: A. S. Lewis, P. A. Parrilo, M. V. Ramana.
1337: The Lax conjecture is true. Proc. Amer. Math. Soc. 133(9):2495--2499, 2005.
1338:
1339: \bibitem{meyerbrandis}
1340: T. Meyer-Brandis. Ber\"uhrungssysteme und symmetrische Darstellungen
1341: ebener Kurven. Diplomarbeit, Univ. Mainz, 1998.
1342:
1343: \bibitem{nn}
1344: Y. Nesterov, A. Nemirovskii. Interior point polynomial algorithms in
1345: convex programming, SIAM, 1994.
1346:
1347: \bibitem{piontkowski}
1348: J. Piontkowski.
1349: Linear symmetric determinantal hypersurfaces.
1350: Michigan Math. J. 54:117-146, 2006.
1351:
1352: \bibitem{polyx}
1353: PolyX, Ltd.
1354: The Polynomial Toolbox for Matlab. 1998.
1355:
1356: \bibitem{renegar}
1357: J. Renegar.
1358: Hyperbolic programs and their derivative relaxations.
1359: Foundations Comput. Math. 6(1):59--79, 2006.
1360:
1361: \bibitem{room}
1362: T. G. Room.
1363: The geometry of determinantal loci. Cambridge Univ. Press, 1938.
1364:
1365: \bibitem{sendra}
1366: J. R. Sendra, F. Winkler. Symbolic parametrization of curves.
1367: J. Symbolic Comput. 12(6):607--631, 1991.
1368:
1369: \bibitem{szilagyi}
1370: I. Szil\'agyi.
1371: Symbolic-numeric techniques for cubic surfaces. PhD Thesis,
1372: RISC, Univ. Linz, Austria, 2005.
1373:
1374: \bibitem{taussky}
1375: O. Taussky. Nonsingular cubic curves as determinantal loci.
1376: J. Math. Phys. Sci. 21(6):665--678, 1987.
1377:
1378: \bibitem{tr}
1379: H. L. Trentelman and P. Rapisarda.
1380: New algorithms for polynomial J-spectral factorization.
1381: Math. Control Signals Systems, 12:24--61, 1999.
1382:
1383: \bibitem{tyrty}
1384: E. E. Tyrtyshnikov. How bad are Hankel matrices ?
1385: Numer. Math. 67:261--269, 1994.
1386:
1387: \bibitem{vanhoeij}
1388: M. van Hoeij. Rational parametrization of curves using canonical
1389: divisors. J. Symbolic Comput. 23:209--227, 1997.
1390:
1391: \bibitem{vinnikov86}
1392: V. Vinnikov.
1393: Self-adjoint determinantal representations of real irreducible
1394: cubics. In: Operator Theory - Advances and Applications,
1395: Vol. 19, Birkh\"auser, 1986.
1396:
1397: \bibitem{vinnikov93}
1398: V. Vinnikov.
1399: Self-adjoint determinantal representations of real plane curves.
1400: Math. Annalen, 296:453--479, 1993.
1401:
1402: \bibitem{willems}
1403: J. C. Willems.
1404: Least squares stationary optimal control and the algebraic
1405: Riccati equation. IEEE Trans. Autom. Control, 16(6):621--634, 1971.
1406:
1407: \bibitem{yakubovich}
1408: V. A. Yakubovich.
1409: Factorization of symmetric matrix polynomials.
1410: Dokl. Acad. Nauk. SSSR, 194(3):1261--1264, 1970.
1411:
1412: \bibitem{zh}
1413: J. C. Z\'u\~niga, D. Henrion.
1414: A Toeplitz algorithm for polynomial J-spectral factorization.
1415: Automatica, 42(7):1085--1093, 2006.
1416:
1417: \end{thebibliography}
1418:
1419: \end{document}
1420: