1: \documentclass[11pt]{amsart}
2: \usepackage{amssymb,amsmath,amscd,latexsym,epsfig,color,hyperref,graphics}
3: \usepackage[all,cmtip]{xy}
4:
5: \newtheorem{theorem}{Theorem}[section]
6: \newtheorem{proposition}[theorem]{Proposition}
7: \newtheorem{corollary}[theorem]{Corollary}
8: \newtheorem{conjecture}[theorem]{Conjecture}
9: \newtheorem{lemma}[theorem]{Lemma}
10: \newtheorem{property}[theorem]{Property}
11: \newtheorem{problem}[theorem]{Problem}
12: \newtheorem{question}[theorem]{Question}
13:
14: \theoremstyle{definition}
15: \newtheorem{definition}[theorem]{Definition}
16: \newtheorem{remark}[theorem]{Remark}
17: \newtheorem{example}[theorem]{Example}
18: \newtheorem{algorithm}[theorem]{Algorithm}
19:
20: \def\NN{\ensuremath{\mathbb{N}}}
21: \def\ZZ{\ensuremath{\mathbb{Z}}}
22: \def\QQ{\ensuremath{\mathbb{Q}}}
23: \def\RR{\ensuremath{\mathbb{R}}}
24: \newcommand{\CC}{{\mathbb C}}
25: \newcommand{\PP}{{\mathbb P}}
26:
27: \newcommand{\V}{{\mathcal V}}
28: \newcommand{\B}{{\mathcal B}}
29: \newcommand{\N}{{\mathcal N}}
30: \def\I{\ensuremath{\mathcal{I}}}
31: \def\F{\ensuremath{\mathcal{F}}}
32: \def\G{\ensuremath{\mathcal{G}}}
33: \def\K{\ensuremath{\mathcal{K}}}
34: \def\A{\ensuremath{\mathcal{A}}}
35: \def\C{\ensuremath{\mathcal{C}}}
36: \def\D{\ensuremath{\mathcal{D}}}
37: \def\E{\ensuremath{\mathcal{E}}}
38: \def\U{\ensuremath{\mathcal{U}}}
39: \def\L{\ensuremath{\mathcal{L}}}
40:
41: \def\aa{\ensuremath{{\bf{a}}}}
42: \def\bb{\ensuremath{{\bf{b}}}}
43: \def\cc{\ensuremath{{\bf{c}}}}
44: \def\dd{\ensuremath{{\bf{d}}}}
45: \def\ee{\ensuremath{{\bf{e}}}}
46: \def\ff{\ensuremath{{\bf{f}}}}
47: \def\hh{\ensuremath{{\bf{h}}}}
48: \def\kk{\ensuremath{{\bf{k}}}}
49: \def\mm{\ensuremath{{\bf{m}}}}
50: \def\pp{\ensuremath{{\bf{p}}}}
51: \def\qq{\ensuremath{{\bf{q}}}}
52: \def\rr{\ensuremath{{\bf{r}}}}
53: \def\ss{\ensuremath{{\bf{s}}}}
54: \def\uu{\ensuremath{{\bf{u}}}}
55: \def\vv{\ensuremath{{\bf{v}}}}
56: \def\ww{\ensuremath{{\bf{w}}}}
57: \def\xx{\ensuremath{{\bf{x}}}}
58: \def\yy{\ensuremath{{\bf{y}}}}
59: \def\zz{\ensuremath{{\bf{z}}}}
60:
61: \def\ica{\ensuremath{{I_{{\mathcal C}_{\mathcal A}}}}}
62: \def\ia{\ensuremath{{{I_{\mathcal A}}}}}
63: \def\iastar{\ensuremath{{{I_{\mathcal A^\ast}}}}}
64: \def\init{\ensuremath{{{\textup{in}_{\prec}}}}}
65: \def\inomega{\ensuremath{\textup{in}_{\omega}}}
66: \def\inomegaprime{\ensuremath{\textup{in}_{\omega^\prime}}}
67: \def\inomegaone{\ensuremath{\textup{in}_{\omega_1}}}
68: \def\inomegatwo{\ensuremath{\textup{in}_{\omega_2}}}
69: \def\inu{\ensuremath{\textup{in}_{\u}}}
70: \def\inv{\ensuremath{\textup{in}_{\v}}}
71: \def\tin{\textup{in}}
72: \def\la{\ensuremath{{{\mathcal{L}_{\mathcal A}}}}}
73: \def\na{\ensuremath{{{\mathbb{N} A}}}}
74: \def\xa{\ensuremath{{{X_{\mathcal A}}}}}
75: \def\hia{\ensuremath{{{\mathbf{H}_{\ia}}}}}
76: \def\hica{\ensuremath{{{\mathbf{H}_{\ica}}}}}
77: \def\HG{\ensuremath{\textup{HG}}}
78: \def\GF{\ensuremath{\textup{GF}}}
79: \def\tif{\textup{if}}
80:
81: \def\mult{\ensuremath{\textup{mult}}}
82: \def\rank{\ensuremath{\textup{rank}}}
83: \def\max{\ensuremath{\textup{max}}}
84: \def\min{\ensuremath{\textup{min}}}
85: \def\supp{\ensuremath{\textup{supp}}}
86: \def\cone{\ensuremath{\textup{cone}}}
87: \def\conv{\ensuremath{\textup{conv}}}
88: \def\cl{\ensuremath{\textup{cl}}}
89: \def\deg{\ensuremath{\textup{deg}}}
90: \def\ker{\ensuremath{\textup{ker}}}
91: \def\ass{\ensuremath{\textup{Ass}}}
92: \def\top{\ensuremath{\textup{top}}}
93: \def\pos{\ensuremath{\textup{pos}}}
94: \newcommand{\iso}{\simeq}
95: \newcommand{\bigo}{{\mathcal O}}
96: \newcommand{\todo}[1]{\vspace{5 mm}\par \noindent \marginpar{\textsc{ToDo}}
97: \framebox{\begin{minipage}[c]{0.95 \textwidth} \tt #1
98: \end{minipage}}\vspace{5 mm}\par}
99:
100:
101: \def\proof{\smallskip\noindent {\it Proof: \ }}
102: \def\endproof{\hfill$\square$\medskip}
103:
104:
105: \title{Theta Bodies for Polynomial Ideals} \author{Jo{\~a}o Gouveia}
106: \address{Department of Mathematics, University of Washington, Box
107: 354350, Seattle, WA 98195, USA, and CMUC, Department of Mathematics,
108: University of Coimbra, 3001-454 Coimbra, Portugal}
109: \email{jgouveia@math.washington.edu} \author{Pablo A. Parrilo}
110: \address{Department of Electrical Engineering and Computer Science,
111: Laboratory for Information and Decision Systems, Massachusetts
112: Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA
113: 02139-4307, USA} \email{parrilo@mit.edu} \author{Rekha R. Thomas}
114: \address{Department of Mathematics, University of Washington, Box
115: 354350, Seattle, WA 98195, USA} \email{thomas@math.washington.edu}
116: \thanks{All authors were partially supported by the NSF Focused
117: Research Group grant (DMS-0757371, DMS-0757207). Gouveia was also
118: supported by Funda{\c c}{\~ a}o para a Ci{\^ e}ncia e Tecnologia,
119: and Thomas by the Robert R. and Elaine K. Phelps Endowed
120: Professorship.} \date{\today}
121:
122: \begin{document}
123:
124: \begin{abstract}
125: Inspired by a question of Lov{\'a}sz, we introduce a hierarchy of
126: nested semidefinite relaxations of the convex hull of real solutions
127: to an arbitrary polynomial ideal, called theta bodies of the ideal.
128: These relaxations generalize Lov{\'a}sz's construction of the theta
129: body of a graph.
130: % We prove that theta bodies are, up to closure, a
131: % modified version of Lasserre's relaxations for real solutions to
132: % ideals, and that they can be computed explicitly using combinatorial
133: % moment matrices.
134: We establish a relationship between theta bodies and Lasserre's
135: relaxations for real varieties which allows, in many cases, for
136: theta bodies to be expressed as feasible regions of semidefinite
137: programs. Examples from combinatorial optimization are given.
138: Lov{\'a}sz asked to characterize ideals for which the first theta
139: body equals the closure of the convex hull of its real variety. We
140: answer this question for vanishing ideals of finite point sets via
141: several equivalent characterizations. We also give a geometric
142: description of the first theta body for all ideals.
143: \end{abstract}
144: \maketitle
145:
146: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
147: \section{Introduction}
148:
149: A central concern in optimization is to understand $\conv(S)$, the
150: convex hull of the set of feasible solutions $S$, to a given problem.
151: In many instances, the set of feasible solutions to an optimization
152: problem is the set of real solutions to a polynomial system: $f_1(\xx)
153: = f_2(\xx) = \cdots = f_m(\xx) = 0$, where $f_1, \ldots, f_m \in
154: \RR[\xx] := \RR[x_1,\ldots,x_n]$. This set is the {\em real variety},
155: $\V_\RR(I)$, of the {\em ideal} $I$ in $\RR[\xx]$ generated by $f_1,
156: \ldots, f_m$, and it is often necessary to compute or represent
157: $\conv(\V_\RR(I))$ exactly or at least approximately.
158:
159: Recall that $\cl(\conv(\V_\RR(I)))$, the closure of
160: $\conv(\V_\RR(I))$, is cut out by the inequalities $f(\xx) \geq 0$ as
161: $f$ runs over all linear polynomials that are non-negative on
162: $\V_{\RR}(I)$. (Call $f \in \RR[\xx]$ a {\em linear} polynomial if it
163: is affine linear of the form $f=a_0 + \sum_{i=1}^{n} a_i x_i$.) A
164: classical certificate for the non-negativity of a polynomial $f$ on
165: $\V_\RR(I)$ is the existence of a {\em sum of squares} ({\em sos})
166: polynomial $\sum_{j=1}^{t} h_j^2$ that is congruent to $f$ mod $I$
167: (i.e., $f-\sum_{j=1}^{t} h_j^2 \in I$), written as $f \equiv
168: \sum_{j=1}^{t} h_j^2$ mod $I$. If this is the case, we say that $f$
169: {\em is sos mod} $I$. Hence a natural relaxation of
170: $\cl(\conv(\V_\RR(I)))$ is the closed convex set:
171: \begin{equation} \label{eqn:sos relaxation} \{ \xx \in \RR^n \,:\,
172: f(\xx) \geq 0 \,\,\forall\,\,f \textup{ linear and sos mod } I \}.
173: \end{equation}
174: Depending on $I$, (\ref{eqn:sos relaxation}) may be strictly larger
175: than $\cl(\conv(\V_\RR(I)))$ since there may be polynomials that are
176: non-negative on $\V_\RR(I)$ but not sos mod $I$. However, in many
177: interesting cases, (\ref{eqn:sos relaxation}) will equal
178: $\cl(\conv(\V_\RR(I)))$. By bounding the degree of the $h_j$'s that
179: appear in the sos representations, and gradually increasing this
180: bound, we obtain a hierarchy of relaxations to
181: $\cl(\conv(\V_\RR(I)))$. In \cite{Lovasz}, Lov{\'a}sz asked a question
182: that leads to the study of this hierarchy. To explain it, we first
183: introduce some definitions.
184:
185: \begin{definition} \label{def:perfect ideals} Let $f$ be a polynomial
186: in $\RR[\xx]$, $I$ be an ideal in $\RR[\xx]$ with real variety
187: $\V_{\RR}(I) := \{\ss \in \RR^n \,:\, f(\ss) = 0 \,\,\forall \,\,f
188: \in I \}$, and let $\RR[\xx]_k$ denote the set of polynomials in
189: $\RR[\xx]$ of degree at most $k$.
190: \begin{enumerate}
191: \item The polynomial $f$ is {\bf $k$-sos}
192: mod $I$ if there exists $h_1,\ldots,h_t \in \RR[\xx]_k$ for some $t$
193: such that $f \equiv \sum_{j=1}^t h_j^2 \,\,\textup{mod}\,\,I$.
194:
195: \item The ideal $I$ is {\bf $k$-sos} if {\em every} polynomial that is
196: non-negative on $\V_\RR(I)$ is $k$-sos mod $I$. If every polynomial
197: of degree at most $d$ that is non-negative on $\V_{\RR}(I)$ is $k$-sos mod
198: $I$, we say that $I$ is {\bf $(d,k)$-sos}.
199: \end{enumerate}
200: \end{definition}
201:
202: \begin{example} \label{ex:runningex} Consider the principal ideal $I =
203: \langle x_1^2x_2 - 1 \rangle \subset \RR[x_1,x_2]$. Then
204: $\textup{conv}(\V_{\RR}(I)) = \{ (s_1,s_2) \in \RR^2 \,:\, s_2 > 0
205: \}$, and any linear polynomial that is non-negative over
206: $\V_{\RR}(I)$ is of the form $\alpha x_2 + \beta$, where $\alpha,
207: \beta \geq 0$. Since $\alpha x_2 + \beta \equiv (\sqrt{\alpha}
208: x_1x_2)^2 + (\sqrt{\beta})^2$ mod $I$, $I$ is $(1,2)$-sos. Check
209: that $x_2$ is not $1$-sos mod $I$ and so, $I$ is not $(1,1)$-sos.
210: \end{example}
211:
212: In \cite{Lovasz}, Lov{\'a}sz asked the following question.
213:
214: \begin{problem} \cite[Problem 8.3]{Lovasz} \label{prob:lovasz} Which
215: ideals in $\RR[\xx]$ are $(1,1)$-sos? How about
216: $(1,k)$-sos?
217: \end{problem}
218:
219: The geometry behind the above algebraic question leads to a natural
220: hierarchy of relaxations of $\textup{conv}(\V_{\RR}(I))$ which we now
221: introduce. The name comes from earlier work of Lov{\'a}sz and will be
222: explained in Section~\ref{sec:examples}.
223:
224: \begin{definition} \label{def:theta}
225: \begin{enumerate}
226: \item For a positive integer $k$, the $k$-th {\bf theta body} of an
227: ideal $I \subseteq \RR[\xx]$ is
228: $$\textup{TH}_k(I) := \{ \xx \in \RR^n \,:\,
229: f(\xx) \geq 0 \,\,\textup{for every linear}
230: \,\,f\,\,\textup{that is $k$-sos mod} \,\,I \}.$$
231: \item An ideal $I \subseteq \RR[\xx]$ is {\bf $\textup{TH}_k$-exact}
232: if $\textup{TH}_k(I)$ equals $\cl(\textup{conv}(\V_{\RR}(I)))$.
233: \item The {\bf theta-rank} of $I$ is the smallest $k$ for which $I$ is
234: $\textup{TH}_k$-exact.
235: \end{enumerate}
236: \end{definition}
237:
238:
239: By definition, $\textup{TH}_1(I) \supseteq \textup{TH}_2(I) \supseteq
240: \cdots \supseteq \textup{conv}(\V_{\RR}(I))$. As seen in
241: Example~\ref{ex:runningex}, $\textup{conv}(\V_{\RR}(I))$ may not be
242: closed while the theta bodies are. Therefore, the theta-body sequence
243: of $I$ can converge, if at all, only to
244: $\textup{cl}(\conv(\V_\RR(I)))$.
245: % If an ideal $I$ is $\textup{TH}_k$-exact, then linear optimization
246: % over $\V_\RR(I)$ reduces to linear optimization over
247: % $\textup{TH}_k(I)$ which, as we will see in Section~\ref{sec:theta
248: % bodies}, can be represented using {\em semidefinite
249: % programming}. Therefore,
250:
251: A natural question at this point is whether
252: the algebraic notion of an ideal being $(1,k)$-sos is equivalent to
253: the geometric notion of being $\textup{TH}_k$-exact.
254:
255: \begin{lemma} \label{lem:1ksos implies thetakexact} If an ideal $I
256: \subseteq \RR[\xx]$ is $(1,k)$-sos then it is $\textup{TH}_k$-exact.
257: \end{lemma}
258:
259: \begin{proof}
260: Let $I$ be $(1,k)$-sos and $\ss \in \RR^n$ be not in
261: $\textrm{cl}(\conv(\V_{\RR}(I)))$. By the {\em separation theorem}
262: \cite[Theorem III.1.3]{Barvinok} there exists a linear polynomial
263: $f$, non-negative over $\textrm{cl}(\conv(\V_{\RR}(I)))$, such that
264: $f(\ss)<0$. However, since $I$ is $(1,k)$-sos, $f$ is $k$-sos mod
265: $I$ and so $\ss \not \in \textrm{TH}_k(I)$. Hence $\textup{TH}_k(I)
266: \subseteq \textrm{cl}(\conv(\V_{\RR}(I)))$ and so,
267: $\textup{TH}_k(I)$ equals $\textrm{cl}(\conv(\V_{\RR}(I)))$.
268: % If $I$ is $(1,k)$-sos then any linear $f$
269: % non-negative on $\V_\RR(I)$ (hence on
270: % $\textup{cl}(\textup{conv}(\V_{\RR}(I)))$) is $k$-sos mod $I$
271: % which implies that $\textup{TH}_k(I) \subseteq
272: % \textup{cl}(\textup{conv}(\V_{\RR}(I)))$. The reverse
273: % inclusion is always true, and so
274: % $\textup{cl}(\textup{conv}(\V_{\RR}(I))) = \textup{TH}_k(I)$.
275: \end{proof}
276:
277: Interestingly, the converse of Lemma~\ref{lem:1ksos implies
278: thetakexact} is false in general.
279:
280: \begin{example} \label{ex:conversefalse} Consider $I = \langle x^2
281: \rangle \subset \RR[x]$ with $\V_{\RR}(I) = \{0\} \subset \RR$. All
282: linear polynomials that are non-negative on $\V_\RR(I)$ are of the
283: form $\pm a^2 x + b^2$ for some $a,b \in \RR$. If $b \neq 0$, then
284: $(\pm a^2 x + b^2) \equiv (\frac{a^2}{2b} x \pm b)^2$ mod
285: $I$. However, $\pm x$ is not a sum of squares mod $I$, and hence $I$
286: is not $(1,k)$-sos for any $k$. On the other hand, $I$ is
287: $\textup{TH}_1$-exact since $\textup{conv}(\V_{\RR}(I)) = \{0\}$ is
288: cut out by the infinitely many linear inequalities $\pm x + b^2 \geq
289: 0$ as $b$ varies over $b \neq 0$.
290: \end{example}
291:
292:
293: \begin{definition} \label{def:ideal defs}
294: Let $I$ be an ideal in $\RR[\xx]$. Then $I$ is
295: \begin{enumerate}
296: \item {\bf radical} if it equals its {\bf radical ideal}
297: $$\sqrt{I} := \{ f \in \RR[\xx] \,:\, f^m \in I, \,\,m \in \NN
298: \backslash \{0\} \},$$
299: \item {\bf real radical} if it equals its {\bf real radical ideal}
300: $$\sqrt[\RR]{I} := \{ f \in \RR[\xx] \,:\, f^{2m} + g_1^2 + \cdots +
301: g_t^2 \in I, \,\, m \in \NN \backslash \{0\}, \,\, g_1, \ldots,
302: g_t \in \RR[\xx] \},$$
303: \item and {\bf zero-dimensional} if its {\bf complex variety}
304: $\V_{\CC}(I) := \{\xx \in \CC^n \,:\, f(\xx) = 0 \,\,\forall \,\, f
305: \in I \}$ is finite.
306: \end{enumerate}
307: \end{definition}
308:
309: Recall that given a set $S \subseteq \RR^n$, its {\em vanishing ideal}
310: in $\RR[\xx]$ is the ideal $\I(S) := \{ f \in \RR[\xx] \,:\, f(\ss)=0
311: \,\,\forall\,\, \ss \in S \}$. {\em Hilbert's Nullstellensatz} states
312: that for an ideal $I \subseteq \RR[\xx]$, $\sqrt{I} = \I(\V_{\CC}(I))$
313: and the {\em Real Nullstellensatz} states that $\sqrt[\RR]{I} =
314: \I(\V_{\RR}(I))$. Hence, $I \subseteq \sqrt{I} \subseteq
315: \sqrt[\RR]{I}$, and if $I$ is real radical then it is also
316: radical. See for example, \cite[Appendix 2]{MarshallBook}, for these
317: notions.
318:
319: We will prove in Section~\ref{sec:theta bodies} that the converse of
320: Lemma~\ref{lem:1ksos implies thetakexact} holds for real radical
321: ideals. These ideals occur frequently in applications and for them,
322: Problem~\ref{prob:lovasz} is asking when $I$ is $\textup{TH}_1$-exact,
323: or more generally, $\textup{TH}_k$-exact.
324:
325: \vspace{.2cm}
326:
327: \noindent{\bf Contents of this paper.}
328: Recall that a {\em semidefinite program}
329: (SDP) is an optimization problem in the space of real symmetric
330: matrices of the form:
331: \begin{equation} \label{sdp}
332: \max \,\,\left\{ \cc^t \xx \,:\, A_0 + \sum_{i=0}^m A_i x_i
333: \succeq 0 \right \},
334: \end{equation}
335: where $\cc \in \RR^m$ and the $A_j$'s are real symmetric matrices. The
336: notation $A \succeq 0$ implies that $A$ is {\em positive
337: semidefinite}. SDPs generalize linear programs and can be solved
338: efficiently \cite{BoydVandenberghe}. In Section~\ref{sec:theta
339: bodies} we prove that under a certain technical hypothesis
340: (satisfied by real radical ideals for instance), the theta body
341: sequence of an ideal $I$ is a modified version of a hierarchy of
342: relaxations for the convex hull of a basic semialgebraic set, due to
343: Lasserre \cite{Lasserre1, Lasserre2}. In this case, each theta body
344: is the closure of the projection of a {\em spectrahedron} (feasible
345: region of a SDP), and an explicit representation is possible using the
346: {\em combinatorial moment matrices} introduced by Laurent
347: \cite{Laurent}. When $I$ is a real radical ideal, we further prove
348: that $I$ is $(1,k)$-sos if and only if $I$ is $\textup{TH}_k$-exact
349: which impacts later sections.
350:
351: In Section~\ref{sec:examples} we illustrate the theta body sequence
352: for the maximum stable set and maximum cut problems in a graph which
353: are two very well-studied problems from combinatorial
354: optimization. The stable set problem motivated
355: Problem~\ref{prob:lovasz}. We explain this connection in detail in
356: Section~\ref{sec:examples}.
357:
358: In Section~\ref{sec:structure} we solve Problem~\ref{prob:lovasz} for
359: vanishing ideals of finite point sets in $\RR^n$. This situation
360: arises often in applications and is the typical set up in
361: combinatorial optimization. Several corollaries follow: If $S \subset
362: \RR^n$ is finite and its vanishing ideal $\I(S)$ is $(1,1)$-sos then
363: $S$ is affinely equivalent to a subset of $\{0,1\}^n$ and its convex
364: hull can have at most $2^n$ facets. If $S$ is the vertex set of a
365: down-closed $0/1$-polytope in $\RR^n$, then $\I(S)$ is $(1,1)$-sos if
366: and only if $\textup{conv}(S)$ is the stable set polytope of a perfect
367: graph. Families of finite sets in growing dimension with $(1,1)$-sos
368: vanishing ideals are exhibited.
369:
370: In Section~\ref{sec:structure arbitrary S}, we give an intrinsic
371: description of the first theta body, $\textup{TH}_1(I)$, of an
372: arbitrary polynomial ideal $I$ in terms of the convex quadrics in $I$.
373: This leads to non-trivial examples of $\textup{TH}_1$-exact ideals
374: with arbitrarily high-dimensional real varieties and reveals the
375: algebraic-geometric structure of $\textup{TH}_1(I)$. Analogous
376: descriptions for higher theta bodies remain open.
377:
378: \begin{remark} In \cite{Lasserre3}, Lasserre introduced the {\em
379: Schm{\"u}dgen Bounded Degree Representation} (S-BDR) and the {\em
380: Putinar-Prestel Bounded Degree Representation} (PP-BDR) properties
381: of a compact basic semialgebraic set $K = \{\xx \,:\, g_1(\xx) \geq
382: 0,\ldots,g_m(\xx) \geq 0\}$ (where $g_i \in \RR[\xx]$),
383: defined as follows:
384: \begin{itemize}
385: \item $K$ has the S-BDR property if there exists a positive integer
386: $k$ such that {\em almost all} linear $f$ that are positive over $K$
387: has a representation as $f = \sum_{J \subseteq [m]} \sigma_J g_J$
388: where $\sigma_J$ are sos, $g_J := \prod_{j \in J} g_j$ and the
389: degree of $\sigma_J g_J$ is at most $2k$ for all $J \subseteq [m] :=
390: \{1,\ldots,m\}$.
391: \item $K$ has the PP-BDR property if there exists a positive integer
392: $k$ such that {\em almost all} linear $f$ that are positive over $K$
393: has a representation as $f = \sum_{j=0}^{m} \sigma_j g_j$ where
394: $\sigma_j$ are sos, $g_0 := 1$ and the degree of $\sigma_j g_j$ is
395: at most $2k$ for $j=0,\ldots,m$.
396: \end{itemize}
397: Call the smallest such $k$ the S-BDR (respectively, PP-BDR) rank of
398: $K$. Here ``almost all'' means all except a set of Lebesgue measure
399: zero. Note that the PP-BDR property implies that S-BDR property.
400:
401: For an ideal $I=\left< f_1,\ldots,f_m \right> \subset \RR[\xx]$,
402: $V_\RR(I)$ is the, possibly non-compact, basic semialgebraic set
403: $\{\xx \in \RR^n \,:\, \pm f_1(\xx) \geq 0, \ldots, \pm f_m(\xx) \geq
404: 0 \}$. When $\V_\RR(I)$ is compact, its PP-BDR property is closely
405: related to the $(1,k)$-sos and $\textup{TH}_k$-exact properties of
406: $I$. However, these notions are not exactly comparable since the
407: PP-BDR rank of $\V_\RR(I)$ depends on the choice of generators of $I$,
408: and only the linear polynomials that are positive (as opposed to
409: non-negative) over $\V_\RR(I)$. Regardless, note that if $\V_{\RR}(I)$
410: has PP-BDR rank $k$, then $I$ has theta-rank at most $k$.
411: \end{remark}
412:
413: {{\bf Acknowledgments}. We thank Monique Laurent and Ting Kei Pong for
414: several useful inputs to this paper. We also thank the referees for
415: their many constructive comments that helped the organization of the
416: paper.}
417:
418: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
419: \section{Theta Bodies} \label{sec:theta bodies}
420:
421: In Definition~\ref{def:theta} we introduced the $k$-th theta body of a
422: polynomial ideal $I \subseteq \RR[\xx]$ and observed that these bodies
423: create a nested sequence of closed convex relaxations of
424: $\textup{conv}(\V_{\RR}(I))$ with $\textup{TH}_k(I) \supseteq
425: \textup{TH}_{k+1}(I) \supseteq \textup{conv}(\V_{\RR}(I))$. Lasserre
426: \cite{Lasserre1} and Parrilo \cite{Parrilo:phd,Parrilo:spr} have
427: independently introduced hierarchies of semidefinite relaxations for
428: polynomial optimization over basic semialgebraic sets in $\RR^n$ using
429: results from real algebraic geometry and the theory of moments. We
430: first examine the connection between the theta bodies of an ideal $I$
431: and Lasserre's relaxations for $\textup{conv}(\V_{\RR}(I))$.
432: % We adopt the point of view in
433: % \cite{MarshallBook} of Lasserre's method to prove our results.
434: % Proposition~\ref{prop:kperfect} shows that if $I$ is real radical then
435: % being $(1,k)$-sos is equivalent to being $\textup{TH}_k$-exact.
436:
437: % Using results in the literature we point out several situations in
438: % which the theta body sequence of an ideal $I$ is guaranteed to
439: % converge to $\textup{cl}(\textup{conv}(\V_{\RR}(I)))$. We also show
440: % how theta bodies can be handled computationally, by directly computing
441: % SOS decompositions in the quotient ring $\RR[\xx]/I$
442: % \cite{SOSstructbook}, or dually, by using the combinatorial moment
443: % matrices from \cite{Laurent}.
444:
445: \subsection{Lasserre's hierarchy and theta bodies}
446:
447: \begin{definition}
448: Let $I$ be an ideal in
449: $\RR[\xx]$. The {\bf quadratic module} of $I$ is
450: $$\mathcal{M}(I):=\left\{s + I \,:\, s
451: \textrm{ is sos in} \,\,\RR[\xx] \right \}.$$
452: The {\bf $k$-th truncation} of $\mathcal{M}(I)$ is
453: $$\mathcal{M}_{k}(I):=\left\{s + I \,:\, s
454: \textrm{ is } k\textrm{-sos} \right\}.$$
455: \end{definition}
456:
457: Both $\mathcal{M}(I)$ and $\mathcal{M}_{k}(I)$ are cones in the
458: $\RR$-vector space $\RR[\xx]/I$. Let $(\RR[\xx]/I)'$ denote the set
459: of linear functionals on $\RR[\xx]/I$ and $\pi_I$ be the projection map
460: from $(\RR[\xx]/I)'$ to $\RR^n$ defined as
461: $\pi_I(y)=(y(x_1+I),\ldots,y(x_n+I)).$ Also let
462: $\mathcal{M}_{k}(I)^{*} \subseteq (\RR[\xx]/I)'$ denote the {\em
463: dual cone} to $\mathcal{M}_{k}(I)$, the set of all linear functions
464: on $\RR[\xx]/I$ that are non-negative on $\mathcal{M}_{k}(I)$.
465:
466: \begin{definition}
467: For $y \in (\RR[\xx]/I)'$, let $H_y$ be the symmetric bilinear
468: form
469: $$
470: \begin{array}{rccc}
471: H_y : & \RR[\xx]/I \times \RR[\xx]/I & \longrightarrow & \RR\\
472: & (f+I,g+I) & \longmapsto & y(fg+I)
473: \end{array}
474: $$
475: and $H_{y,t}$ be the restriction of $H_y$ to the subspace
476: $\RR[\xx]_{t}/I$.
477: \end{definition}
478:
479: Recall that a symmetric bilinear form $H:V \times V \rightarrow \RR$,
480: where $V$ is a $\RR$-vector space, is positive semidefinite
481: (written as $H \succeq 0$) if $H(v,v) \geq 0$ for all non-zero
482: elements $v \in V$. Given a basis $B$ of $V$, the matrix indexed by
483: the elements of $B$ with $(b_i,b_j)$-entry equal to $H(b_i,b_j)$ is
484: called the {\em matrix representation} of $H$ in the basis $B$. The
485: form $H$ is positive semidefinite if and only if its matrix
486: representation in any basis is positive semidefinite.
487:
488: \begin{lemma} \label{thm:sdpformulation} Let $I \subseteq \RR[\xx]$ be
489: an ideal and $k$ a positive integer. Then
490: $$\mathcal{M}_{k}(I)^{*}=\{y \in (\RR[\xx]/I)' : H_{y,k} \succeq 0\}.$$
491: \end{lemma}
492: \begin{proof}
493: Note that $y \in \mathcal{M}_{k}(I)^{*}$ if and only if $y(s + I)
494: \geq 0$ for all $k$-sos polynomials $s$. By linearity of $y$ this is
495: equivalent to $y(h^2 + I) \geq 0$ for all $h \in \RR[\xx]_k$ which
496: is the definition of $H_{y,k}$ being positive semidefinite.
497: \end{proof}
498:
499: The original Lasserre relaxations in \cite{Lasserre1} approximate
500: $\conv(S)$ for a basic semialgebraic set $S=\{\xx \in \RR^n \,:\,
501: g_i(\xx) \geq 0,\, i=1,\ldots,m\}$ by the sets
502: $$\left \{ (y(x_1),\ldots,y(x_n)) \,:\, y \in \RR[\xx]', \,\, y(1) = 1, \,\,
503: y\left(\sum_{i=0}^{m} s_ig_i \right ) \geq 0 \right \}$$ where $s_i$
504: are sos, $g_0 := 1$ and the degree of $s_ig_i$ is bounded above by
505: some fixed positive integer. When there are equations among the
506: $g_i(\xx) \geq 0$, both Lasserre \cite{Lasserre2} (for $0/1$ point
507: sets) and Laurent \cite{Laurent} (more generally for finite varieties)
508: propose doing computations mod the ideal generated by the polynomials
509: defining the equations, to increase efficiency. We adopt this point of
510: view since in our case, $S = \V_\RR(I)$, is cut out entirely by
511: equations, and work with the following definition of a Lasserre
512: relaxation.
513:
514: \begin{definition} \label{def:lasserre relaxation} Let $I \subseteq
515: \RR[\xx]$ be an ideal, $k$ be a positive integer, and
516: $\mathcal{Y}_1$ be the hyperplane of all functions $y \in
517: (\RR[\xx]/I)'$ such that $y(1+I)=1$. The {\bf $k$-th modified
518: Lasserre relaxation} $Q_{k}(I)$ of $\textup{conv}(\V_{\RR}(I))$ is
519: $$Q_{k}(I):=\pi_I(\mathcal{M}_{k}(I)^* \cap \mathcal{Y}_1).$$
520: \end{definition}
521:
522: While $\mathcal{M}_{k}(I)^* \cap \mathcal{Y}_1$ is always closed,
523: $Q_k(I)$ might not be (see Example~\ref{ex:runningex2}). We first note
524: that $Q_k(I)$ is indeed a relaxation of $\conv(\V_\RR(I))$.
525:
526: \begin{lemma} \label{lem:QkI contains conv}
527: For an ideal $I$ and a positive integer $k$,
528: $\textup{conv}(\V_{\RR}(I)) \subseteq Q_k(I)$.
529: \end{lemma}
530:
531: \begin{proof}
532: For $\ss \in \V_{\RR}(I)$, consider $y^{\ss} \in (\RR[\xx]/I)'$
533: defined as $y^{\ss}(f+I) := f(\ss)$. Then $y^{\ss} \in
534: \mathcal{M}_{k}(I)^*$ and $y^{\ss}(1+I) = 1$. Therefore,
535: $\pi_I(y^{\ss})=\ss \in Q_{k}(I)$, and $\conv(\V_{\RR}(I)) \subseteq
536: Q_{k}(I)$ since $Q_{k}(I)$ is convex.
537: \end{proof}
538:
539: Since $Q_{k+1}(I) \subseteq Q_{k}(I)$, these bodies create a nested
540: sequence of relaxations of $\textup{conv}(\V_{\RR}(I)))$ as
541: intended. Our main goal in this section is to establish a relationship
542: between $Q_k(I)$ and the $k$-th theta body, $\textup{TH}_k(I)$, of the
543: ideal $I$ (cf. Definition \ref{def:theta}). We start by noting the
544: following inclusion.
545:
546: \begin{proposition} \label{prop:basis free formulation} For an ideal
547: $I \subseteq \RR[\xx]$ and a positive integer $k$,
548: $\textup{cl}(Q_{k}(I)) \subseteq \textup{TH}_k(I)$.
549: \end{proposition}
550: \begin{proof}
551: Since $\textup{TH}_k(I)$ is closed, it is enough to show that
552: $Q_{k}(I) \subseteq \textup{TH}_k(I)$. Pick $\pp \in Q_{k}(I)$ and
553: $y \in \mathcal{M}_{k}(I)^* \cap \mathcal{Y}_1 $ such that
554: $\pi_I(y)=\pp$. Let $f = a_0 + \sum_{i=1}^{n} a_i x_i$ and $f+I \in
555: \mathcal{M}_{k}(I)$. Then, $\pp \in \textup{TH}_k(I)$ since
556: $$f(\pp) = f(\pi_I(y))= a_0 y(1+I)+ \sum_{i=1}^{n} a_i y(x_i + I) =
557: y(f+I) \geq 0.$$
558: \end{proof}
559:
560: Theorem~\ref{thm:basis free formulation} will prove that if ${\mathcal
561: M}_k(I)$ is closed, we have the equality $\textup{cl}(Q_{k}(I)) =
562: \textup{TH}_k(I)$.
563:
564: \begin{lemma} \label{lem:new lemma} Let $I \subseteq \RR[\xx]$ be an
565: ideal and $k$ be a positive integer. If $f \in \RR[\xx]_1$ is
566: non-negative over $Q_k(I)$, then $f+I \in \textup{cl}({\mathcal
567: M}_k(I))$.
568: \end{lemma}
569:
570: \begin{proof} Suppose $f \in \RR[\xx]_1$ is non-negative over $Q_k(I)$
571: and $f+I \not \in \textup{cl}({\mathcal M}_k(I))$. Then by the
572: separation theorem, there exists $y \in {\mathcal M}_k(I)^*$ such
573: that $y(f+I) < 0$. Since $(f+r+I)^2 = (f+r)^2+I$ lies in ${\mathcal
574: M}_k(I)$ for any real number $r$, $y \in {\mathcal M}_k(I)^*$ and
575: $y$ is linear, we get
576: $$0 \leq y((f+r+I)^2) = y(f^2+I) + 2ry(f+I) + r^2y(1+I)$$
577: which implies that $y(1+I) > 0$ since $y(f+I) \neq 0$. Scaling $y$
578: such that $y(1+I) = 1$, we have that $y \in {\mathcal M}_k(I)^* \cap
579: {\mathcal Y}_1$. This implies that $\pi_I(y) \in Q_k(I)$ and
580: therefore, by hypothesis, $f(\pi_I(y)) \geq 0$. However, since $f \in
581: \RR[\xx]_1$ and $y$ is linear, we also get $f(\pi_I(y)) = y(f+I) < 0$
582: which is a contradiction.
583: \end{proof}
584:
585: \begin{theorem} \label{thm:basis free formulation} Let $I \subseteq
586: \RR[\xx]$ be an ideal. For a positive integer $k$, if ${\mathcal
587: M}_k(I)$ is closed, then $\textup{cl}(Q_{k}(I)) =
588: \textup{TH}_k(I)$.
589: \end{theorem}
590:
591: \begin{proof}
592: By Proposition~\ref{prop:basis free formulation}, we need to prove
593: that when ${\mathcal M}_k(I)$ is closed, $\textup{TH}_k(I) \subseteq
594: \textup{cl}(Q_{k}(I))$. Suppose $\pp \not \in
595: \textup{cl}(Q_{k}(I))$. By the separation theorem, there exists $f
596: \in \RR[\xx]_1$ non-negative on $\textup{cl}(Q_{k}(I))$ with $f(\pp)
597: < 0$. By Lemma~\ref{lem:new lemma}, $f+I \in {\mathcal M}_k(I)$
598: since ${\mathcal M}_k(I)$ is closed by assumption and hence $f$ is
599: $k$-sos mod $I$. Since $f(\pp) < 0$, $\pp \not \in
600: \textup{TH}_k(I)$.
601: \end{proof}
602:
603:
604: An important class of ideals for which ${\mathcal M}_k(I)$ is closed
605: is the set of real radical ideals which are the focus of
606: Sections~\ref{sec:examples} and ~\ref{sec:structure}. We now derive
607: various corollaries to Theorem~\ref{thm:basis free formulation} that
608: apply to real radical ideals.
609:
610: \begin{corollary} \label{cor:equality for real radical ideals} If $I
611: \subseteq \RR[\xx]$ is a real radical ideal then
612: $\textup{cl}(Q_{k}(I)) = \textup{TH}_k(I)$.
613: \end{corollary}
614:
615: \begin{proof}
616: By \cite[Prop 2.6]{PowSchei}, if $I$ is real radical, then ${\mathcal
617: M}_k(I)$ is closed.
618: \end{proof}
619:
620: \begin{lemma} \label{lem:borweinlewis} Let $V$ and $W$ be finite
621: dimensional vector spaces, $H \subseteq W$ be a cone and $A \,:\, V
622: \rightarrow W$ be a linear map such that $A(V) \cap \textup{int}(H)
623: \neq \emptyset$. Then $(A^{-1}H)^* = A'(H^*)$ where $A'$ is the dual
624: operator to $A$. In particular, $A'(H^*)$ is closed in $V'$.
625: \end{lemma}
626:
627: \begin{proof}
628: This follows from Corollary~3.3.13 in \cite{BorweinLewis} by
629: setting $K = V$.
630: \end{proof}
631:
632: \begin{corollary}\label{cor:theta-lasserre}
633: Let $I$ be a real radical ideal in $\RR[\xx]$ and $k$ be a positive
634: integer. If there exists $g \in \RR[\xx]_1$ such that $g+I$ is in
635: the interior of $\mathcal{M}_{k}(I)$ (considered as a subset of
636: $\RR[\xx]_{2k}/I$), then $\textup{TH}_{k}(I)=Q_{k}(I)$.
637: \end{corollary}
638:
639: \begin{proof}
640: By Corollary~\ref{cor:equality for real radical ideals}, it suffices
641: to show that $Q_k(I)$ is closed. Consider $\RR^{n+1}$ with coordinates
642: indexed $0,1,\ldots,n$ and the map
643: $$\tilde{\pi}_I \,:\, (\RR[\xx]/I)' \rightarrow \RR^{n+1}
644: \,\,\textup{such that} \,\,y \mapsto (y(1+I),\pi_I(y)).$$ Then
645: $\tilde{\pi}_I({\mathcal M}_k(I)^* \cap {\mathcal Y}_1) = \{(1,\pp)
646: \,:\, \pp \in Q_k(I) \}$ and so, $Q_k(I)$ will be closed if
647: $\tilde{\pi}_I({\mathcal M}_k(I)^* \cap {\mathcal Y}_1) =
648: \tilde{\pi}_I({\mathcal M}_k(I)^*) \cap \{\pp \in \RR^{n+1} \,:\, p_0
649: = 1 \}$ is closed. Hence, it suffices to show that
650: $\tilde{\pi}_I({\mathcal M}_k(I)^*) \subseteq \RR^{n+1}$ is closed.
651:
652: Now consider the inclusion map $A \,:\, \RR[\xx]_1/I \rightarrow
653: \RR[\xx]_{2k}/I$ and let $M$ denote the cone ${\mathcal M}_k(I)$
654: considered as a subset of $\RR[\xx]_{2k}/I$. By assumption,
655: $A(\RR[\xx]_1/I) \cap \textup{int}(M) = (\RR[\xx]_1/I) \cap
656: \textup{int}(M) \neq \emptyset$ and so by
657: Lemma~\ref{lem:borweinlewis}, $A'(M^*)$ is closed. Let $C :=
658: \{1^*,x_1^*,\ldots,x_n^*\}$ be the canonical basis of
659: $(\RR[\xx]_1/I)'$. Then $A'(\bar{y}) = ({\bar y}(1+I),{\bar
660: y}(x_1+I), \ldots, {\bar y}(x_n+I))$ with respect to $C$. Now note
661: that if $y \in {\mathcal M}_k(I)^*$ then its restriction ${\hat y}$ to
662: $\RR[\xx]_{2k}/I$ belongs to $M^*$ and $A'({\hat y}) =
663: \tilde{\pi}_I(y)$. Therefore, $\tilde{\pi}_I({\mathcal M}_k(I)^*)
664: \subseteq A'(M^*)$. Conversely, if ${\bar y} \in M^*$ and ${\tilde
665: y}$ is any extension to $\RR[\xx]/I$, then ${\tilde y}$ belongs to
666: ${\mathcal M}_k(I)^*$. Since $A'({\bar y}) =
667: \tilde{\pi}_I(\tilde{y})$ we get $A'(M^*) \subseteq
668: \tilde{\pi}_I({\mathcal M}_k(I)^*)$ and so, $A'(M^*) =
669: \tilde{\pi}_I({\mathcal M}_k(I)^*)$ is closed.
670: \end{proof}
671:
672: % If such a $g$ exists then the closure in
673: % \eqref{eq:closure} can be dropped \cite[Corollary
674: % 16.4.2]{Rockafellar}. Then we just have to observe that
675: % $(\mathcal{M}_{k}(I)^{*} + R_1^{\perp}) \cap \mathcal{Y}_1 $ is the
676: % same as $\mathcal{M}_{k}(I)^{*} \cap \mathcal{Y}_1 + R_1^{\perp}$
677: % and the argument used in the proof of the Theorem \ref{thm:basis
678: % free formulation} will yield the intended result.
679:
680: Let $I$ be an ideal and $k$ a positive integer. In
681: Lemma~\ref{lem:1ksos implies thetakexact} we saw that if $I$ is
682: $(1,k)$-sos (i.e., every linear polynomial that is non-negative on
683: $\V_{\RR}(I)$ is $k$-sos mod $I$), then $I$ is $\textup{TH}_k$-exact
684: (i.e., $\textup{TH}_k(I) =
685: \textup{cl}(\textup{conv}(\V_{\RR}(I)))$). Example~\ref{ex:conversefalse}
686: showed that the reverse implication does not always hold.
687:
688: \begin{corollary} \label{cor:real radical equivalence} If an ideal $I
689: \subseteq \RR[\xx]$ is real radical then $I$ is $(1,k)$-sos if and
690: only if $I$ is $\textup{TH}_k$-exact.
691: \end{corollary}
692:
693: \begin{proof}
694: If $f \in \RR[\xx]_1$ is non-negative on $\V_\RR(I)$ and $I$ is
695: $\textup{TH}_k$-exact, then $f$ is non-negative on
696: $\textup{TH}_k(I)$ and hence on $Q_k(I)$. Therefore, by
697: Lemma~\ref{lem:new lemma}, $f \in \textup{cl}({\mathcal M}_k(I))$.
698: Suppose now that $I$ is also real radical. Then ${\mathcal M}_k(I)$
699: is closed and $f+I \in {\mathcal M}_k(I)$, which means that $I$ is
700: $(1,k)$-sos.
701: \end{proof}
702:
703: % \begin{proposition} \label{prop:kperfect} If an ideal $I \subset
704: % \RR[\xx]$ is $\textup{TH}_k$-exact for some positive integer $k$,
705: % and $f \in \RR[\xx]_1$ is non-negative on $\V_{\RR}(I)$, then $f+I
706: % \in \textup{cl}(\mathcal{M}_{k}(I))$.
707: % \end{proposition}
708:
709: % \begin{proof}
710: % Let $f \in \RR[\xx]_1$ be non-negative on $\V_{\RR}(I)$ and suppose
711: % $f+I \not \in \textup{cl}(\mathcal{M}_{k}(I))$. By the separation
712: % theorem, there exists $y \in \mathcal{M}_{k}(I)^*$ such that
713: % $y(f+I)<0$. Given any real number $r$ we then have
714: % $$0 \leq y((f+r+I)^2) = y(f^2+I)+2ry(f+I)+r^2y(1+I).$$
715: % Since $y(f+I)<0$, we must have $y(1+I)>0$. Now scaling $y$ so that
716: % $y(1+I)=1$, we get $0 > y(f+I) = f(\pi_I(y))$ since $f$ is
717: % linear. However, since $y \in \mathcal{M}_{k}(I)^* \cap
718: % \mathcal{Y}_1$, we have $\pi_I(y) \in \textup{TH}_k(I) =
719: % \textup{cl}({\textup{conv}(\V_{\RR}(I))})$, which implies by
720: % hypothesis that $f(\pi_I(y)) \geq 0$. This contradiction implies that
721: % $f+I$ belongs to $\textup{cl}(\mathcal{M}_k(I))$.
722: % \end{proof}
723:
724:
725: % Pick $\pp \in \textup{TH}_k(I)$ (i.e., $f(\pp) \geq 0$ for all
726: % linear $f$ such that $f+I \in \mathcal{M}_{k}(I)$). Choose a basis
727: % $f_1,f_2,\ldots$ of the $\RR$-vector space $I$ such that
728: % $f_1,\ldots,f_t$ is a basis for $I \cap \RR[\xx]_1$. Complete it to
729: % a basis of $\RR[\xx]$ by adding polynomials $h_1,h_2,\ldots$ such
730: % that $f_1,\ldots,f_t,h_1,\ldots,h_s$ is a basis for $\RR[\xx]_1$. We
731: % can then define a linear function $y'$ in $\RR[\xx]'$ by setting
732: % $$y'\left(\sum_j \alpha_j f_j + \sum_i \beta_i h_i)
733: % \right) := \sum_{i=1}^s \beta_i h_i(\pp).$$ Since $y'(I)=\{0\}$ by
734: % definition, we can descend $y'$ to a linear operator $y \in
735: % (\RR[\xx]/I)'$ by simply setting $y(f+I)=y'(f)$ for all $f \in
736: % \RR[\xx]$. For $i=1,\ldots,t$, since $f_i+I=-f_i+I = I \subseteq
737: % \mathcal{M}_{k}(I)$, and $f_i$ is linear, we must have $f_i(\pp)=0$.
738: % This implies that for any other linear polynomial $g$ (which equals
739: % $\sum_{j=1}^t \alpha_{j} f_j + \sum_{i=1}^s \beta_{i} h_i$ for some
740: % constants $\alpha_j$, $\beta_i$), $g(\pp) = \sum_{i=1}^s \beta_{i}
741: % h_i(\pp) = y(g+I)$. In particular, $\pi_I(y) = \pp$, $y(1+I) = 1$,
742: % and if $g$ is linear with $g+I \in \mathcal{M}_{k}(I)$, then $y(g+I)
743: % = g(\pp) \geq 0$. If $R_1:=\RR[\xx]_1 +I$, then what we have so far
744: % is that $y \in \left( R_1 \cap \mathcal{M}_{k}(I)\right)^{*} \cap
745: % \mathcal{Y}_1 $ and $\pi_I(y) = \pp$. To finish the proof, we will
746: % show that $\pi_I(y)$ belongs to the closure of $Q_k(I)$. Let
747: % $R_1^{\perp} \subset (\RR[\xx]/I)'$ be the orthogonal complement of
748: % $R_1$. Note that $\left( R_1 \cap \mathcal{M}_{k}(I)\right)^{*} =
749: % \textup{cl}(\mathcal{M}_{k}(I)^{*} + R_1^{\perp})$ where the closure
750: % is taken in the weak$^*$ topology of $(\RR[\xx]/I)'$. This happens
751: % because both $R_1$ and ${M}_{k}(I)$ are closed, and the dual of the
752: % intersection of two closed cones is the closure of the sum of the
753: % dual cones \cite[Cor.2, p.126]{Schaefer}. Hence,
754: % \begin{equation}\label{eq:closure}
755: % y \in \textup{cl}(\mathcal{M}_{k}(I)^{*} +
756: % R_1^{\perp}) \cap \mathcal{Y}_1.
757: % \end{equation}
758: % Then $y=\lim_{\lambda} (y_{\lambda} + z_{\lambda})$ for some nets
759: % $y_{\lambda} \in \mathcal{M}_{k}(I)^{*}$ and $z_{\lambda} \in
760: % R_1^{\perp}$. Since $z_{\lambda}(1+I) = 0$ for all $\lambda$, we must
761: % have $\lim_\lambda y_{\lambda}(1+I)=y(1+I)=1$ and so
762: % $$y=\lim_{\lambda} \frac{y_{\lambda}}{y_{\lambda}(1+I)} + z_{\lambda}
763: % \in \textup{cl}(\mathcal{M}_{k}(I)^{*}\cap \mathcal{Y}_1 +
764: % R_1^{\perp}).$$
765: % This means that the projection $\pi_I(y)$ must be in the set
766: % $$\pi_I(\textup{cl}(\mathcal{M}_{k}(I)^{*}\cap \mathcal{Y}_1 +
767: % R_1^{\perp})) \subseteq \textup{cl}(\pi_I(\mathcal{M}_{k}(I)^{*}\cap
768: % \mathcal{Y}_1 + R_1^{\perp})),$$ and because $R_1^{\perp}$ lies in the
769: % kernel of $\pi_I$, we conclude that $\pi_I(y)$ lies in
770: % $$\textup{cl}(\pi_I( \mathcal{M}_{k}(I)^{*} \cap \mathcal{Y}_1)) =
771: % \textup{cl}(Q_{k}(I)).$$
772:
773:
774: We close with a brief discussion of ideals for which the theta body
775: sequence is guaranteed to converge (finitely or asymptotically) to
776: $\textup{cl}(\conv(\V_{\RR}(I)))$.
777:
778: \begin{enumerate}
779: \item If $\V_{\RR}(I)$ is finite the results in \cite{LaLauRos} imply
780: that $I$ is $\textup{TH}_k$-exact for some finite $k$. If
781: $\V_\CC(I)$ is finite ($I$ is zero-dimensional), then $k$ can be
782: bounded above by the maximum degree of a linear basis of
783: $\RR[\xx]/I$ \cite{Laurent} (see Section~\ref{subsec:combinatorial
784: matrices}). However, as in $I = \langle x^2 \rangle$, we cannot
785: guarantee that $I$ is $(1,k)$-sos for any $k$, even when $I$ is
786: zero-dimensional. If $I$ is zero-dimensional and radical, then in
787: fact, $I$ is $(1,k)$-sos for finite $k$ with $k \leq
788: |\V_{\CC}(I)|-1$ (see \cite{parrilo}, \cite[Theorem
789: 2.4]{LaurentSosSurvey}). Better bounds are often possible as in
790: Remark~\ref{rem:k parallel translates}. For an ideal $I \subseteq
791: \RR[\xx]$ we summarize the above results in the following table.
792:
793: % \begin{center}
794: % \begin{tikzpicture}
795: % \matrix(m)[matrix of math nodes, row sep=3em col sep=3em, text
796: % height=1.5ex, text depth=0.25ex]
797: % {A&B\\C&D\\};
798: % \path[->,font=scriptsize];
799: % (m-1-1) edge (m-1-2)
800: % (m-1-1) edge (m-2-1)
801: % (m-1-2) edge (m-2-2)
802: % (m-2-1) edge (m-2-2);
803: % \end{tikzpicture}
804: % \end{center}
805:
806:
807: \begin{center}
808: $
809: \xymatrix{
810: \V_\CC(I) \,\,\text{finite} \ar[r] \ar[d]^{\tiny{I = \sqrt{I}}} & \V_\RR(I) \,\,\text{finite} \ar[d] \\
811: I \,\,\, \text{(1,k)-sos} \ar@<-1mm>[r] & I\,\,\,\textup{TH}_k\textup{-exact} \ar@<-1mm>[l]_{\tiny{I=\sqrt[\RR]{I}}}
812: }
813: $
814: \end{center}
815:
816:
817: \item If $\V_{\RR}(I)$ is not finite but is compact, Schm{\"u}dgen's
818: Positivstellensatz \cite[Chapter 3]{MarshallBook} implies that the
819: theta body sequence of $I$ converges (at least asymptotically) to
820: $\textup{cl}(\textup{conv}(\V_{\RR}(I)))$ (i.e.,
821: $\bigcap_{k=1}^{\infty} \textup{TH}_k(I) =
822: \textup{cl}(\textup{conv}(\V_{\RR}(I)))$).
823: % Under strong additional
824: % requirements on smoothness and curvature, results of Helton and Nie
825: % \cite{helton-nie} guarantee that $I$ is $\textup{TH}_k$-exact for
826: % finite $k$.
827:
828: \item If $\V_{\RR}(I)$ is not compact, then the study of the theta
829: body hierarchy becomes harder. Scheiderer \cite[Chapter
830: 2]{MarshallBook} has identified ideals $I$ with $\V_{\RR}(I)$ not
831: necessarily compact, but of dimension at most two, for which every
832: $f \geq 0$ mod $I$ is sos mod $I$. In all these cases, the theta
833: body sequence of $I$ converges to
834: $\textup{cl}(\textup{conv}(\V_{\RR}(I)))$.
835: \end{enumerate}
836:
837:
838: The results of Schm{\"u}dgen and Scheiderer mentioned above fit
839: within a general framework in real algebraic geometry that is
840: concerned with when an arbitrary $f \in \RR[\xx]$ that is positive
841: or non-negative over a basic semi-algebraic set is sos modulo
842: certain algebraic objects defined by the set. We only care about
843: real varieties and whether linear polynomials that are non-negative
844: over them are sos mod their ideals. Therefore, often there are
845: ideals $I$ that are $\textup{TH}_k$-exact or $(1,k)$-sos for which
846: there are non-linear polynomials $f$ such that $f \geq 0$ mod $I$
847: but $f$ is not sos mod $I$. For instance, the proof of
848: Theorem~\ref{thm:intersection} will show that $J_n := \langle
849: \sum_{i=1}^{n} x_i^2 -1 \rangle$ is $(1,1)$-sos for all $n$, but a
850: result of Scheiderer \cite[Theorem 2.6.3]{MarshallBook} implies that
851: when $n \geq 4$, there is always some non-linear $f$ non-negative on
852: $\V_\RR(J_n)$ that is not sos mod $J_n$.
853:
854:
855: \subsection{Combinatorial moment matrices} \label{subsec:combinatorial
856: matrices}
857:
858: To compute theta bodies we must work with the truncated quadratic
859: module $\mathcal{M}_{k}(I)$ which requires computing sums of squares
860: in $\RR[\xx]/I$ as described in \cite{SOSstructbook}, or dually, using
861: the combinatorial moment matrices introduced by Laurent in
862: \cite{Laurent}. We describe the latter viewpoint here as it is more
863: natural for theta bodies.
864:
865: Consider a basis $\B = \{ f_0+I, f_1+I, \ldots \}$ for $\RR[\xx]/I$,
866: and define $\textup{deg}(f_i + I) := \textup{min}_{f-f_i \in I}
867: \textup{deg}\,f$. For a positive integer $k$, let $\B_k := \{ f_l + I
868: \in \B \,:\, \textup{deg}(f_l+I) \leq k \}$, and set $\ff_k := (f_l+I
869: \,:\, f_l+I \in \B_k)$. We may assume that the elements of $\B$ are
870: indexed in order of increasing degree. Let $\lambda^{(g+I)} :=
871: (\lambda_l^{(g+I)})$ be the vector of coordinates of $g+I$ with
872: respect to $\B$. Note that $\lambda^{(g+I)}$ has only finitely many
873: non-zero coordinates.
874:
875: \begin{definition}
876: Let $\yy \in \RR^{\mathcal{B}}$. Then the {\bf combinatorial moment
877: matrix} $M_{\mathcal{B}}(\yy)$ is the (possibly infinite) matrix
878: indexed by $\mathcal{B}$ whose $(i,j)$ entry is
879: $$\lambda^{(f_i f_j +I)} \cdot \yy = \sum \lambda_l^{(f_if_j+I)}
880: y_l.$$
881: The {\bf $k$-th-truncated combinatorial moment matrix}
882: $M_{\B_k}(\yy)$ is the finite (upper left principal) submatrix of
883: $M_{\B}(\yy)$ indexed by $\B_k$.
884: \end{definition}
885:
886: Although only a finite number of the components in
887: $\lambda^{(f_if_j+I)}$ are non-zero, for practical purposes we need to
888: control exactly which indices can be non-zero. One way to do this is
889: by choosing $\B$ such that if $f+I$ has degree $k$ then $f+I \in
890: \textup{span}(\B_k)$. This is true for instance if $\B$ is the set of
891: {\em standard monomials} of a {\em term order} that respects degree
892: \cite{CLO}. If $\B$ has this property then $M_{\B_k}(\yy)$ only
893: depends on the entries of $\yy$ indexed by $\B_{2k}$.
894:
895: \begin{theorem}\label{thm:combinatorial matrices formulation}
896: For each positive integer $k$,
897: $$\textup{proj}_{\RR^{\B_1}}\{ \yy \in \RR^{\B_{2k}} \,:\,
898: M_{\B_k}(\yy) \succeq 0, \, y_0 = 1\} = \ff_1 (Q_{k}(I)),$$ where
899: $y_0$ is the first entry of $\yy \in \RR^{B_{2k}}$,
900: $\textup{proj}_{\RR^{\B_1}}$ is the projection onto the coordinates
901: indexed by $\B_1$, and for $\pp \in \RR^n$, $\ff_1(\pp) :=
902: (f_i(\pp))_{f_i+I \in \B_1}$.
903: \end{theorem}
904: \begin{proof}
905: We may identify $\yy =(y_i) \in \RR^{\B_{2k}}$ with the operator
906: $\bar{y} \in (\RR[\xx]/I)'$ where $\bar{y}(f_i+I)=y_i$ if $f_i+I
907: \in \B_{2k}$ and zero otherwise. Then $M_{\B_k}(\yy)$ is simply the
908: matrix representation of $H_{\bar{y},k}$ in the basis $\B$, since
909: we assumed that if $\deg \,(f_i+I), \deg \,(f_j
910: + I) \leq k$ then $\bar{y}(f_if_j+I)$ depends only on the value of
911: $\bar{y}$ on $\B_{2k}$. Therefore, $\textup{proj}_{\RR^{\B_1}}\{ \yy
912: \in \RR^{\B_{2k}} \,:\, M_{\B_k}(\yy) \succeq 0\}$
913: equals $$\{(\bar{y}(f_i+I))_{\B_1} : \bar{y} \in (\RR[\xx]/I)',
914: H_{\bar{y},k} \succeq 0\}.$$
915: Furthermore, since $f_i$ is linear whenever $f_i+I \in \B_1$,
916: $$(\bar{y}(f_i+I))_{\B_1}= (f_i(\pi_I(\bar{y})))_{\B_1} =:
917: \ff_1(\pi_I(\bar{y}))$$ so by Lemma~\ref{thm:sdpformulation},
918: $\textup{proj}_{\RR^{\B_1}}\{ \yy \in \RR^{\B_{2k}} \,:\,
919: M_{\B_k}(\yy) \succeq 0, \, y_0 = 1\} = \ff_1 (Q_{k}(I)$.
920: \end{proof}
921:
922: \begin{corollary} \label{cor:simple version} Suppose $\B_1 = \{1+I,
923: x_1+I, \ldots, x_n+I\}$ and denote by $y_0,y_1, \ldots, y_n$ the
924: first $n+1$ coordinates of $\yy \in \RR^{\B_{2k}}$, then $$Q_{k}(I)
925: = \{ (y_1, \ldots, y_n) \,:\, \yy \in
926: \RR^{\B_{2k}}\,\textup{with}\,\, M_{\B_k}(\yy) \succeq 0
927: \,\textup{and} \, y_0 = 1\}.$$
928: \end{corollary}
929:
930: By Corollary \ref{cor:simple version}, optimizing a linear function
931: over $Q_k(I)$, hence over $\textup{cl}(Q_k(I))$, is
932: an SDP and can be solved efficiently.
933:
934: \begin{example} \label{ex:runningex2} Consider the ideal $I = \langle
935: x_1^2x_2 - 1 \rangle \subset \RR[x_1,x_2]$ from
936: Example~\ref{ex:runningex} for which $\textup{conv}(\V_{\RR}(I)) =
937: \{(s_1,s_2) \in \RR^2 \,:\, s_2 > 0 \}$ was not closed but $I$ was
938: $\textup{TH}_2$-exact and $(1,2)$-sos. Note that $\B = \bigcup_{k
939: \in \NN} \{x_1^k+I, x_2^k+I, x_1x_2^k+I\}$ is a degree-compatible
940: monomial basis for $\RR[x_1,x_2]/I$ for which $$\B_4 =
941: \{1,x_1,x_2,x_1^2,x_1x_2,x_2^2,x_1x_2^2,x_1^3,x_2^3,x_1x_2^3,x_1^4,x_2^4
942: \}+I.$$ The combinatorial moment matrix $M_{\B_2}(\yy)$ for
943: $\yy=(1,y_1,\ldots, y_{11}) \in \RR^{\B_4}$ is
944: $$ \begin{array}{ll}
945: & \hspace*{.2cm} \begin{array}{cccccc} 1 & \, x_1 & x_2 &
946: x_1^2 & x_1x_2 & \hspace*{-.1cm} x_2^2
947: \end{array} \\
948: \vspace*{-.2cm}
949: & \\
950: \begin{array}{c}
951: 1 \\ x_1 \\ x_2 \\ x_1^2 \\ x_1x_2 \\ x_2^2
952: \end{array} & \hspace*{-.2cm}
953: \left( \begin{array}{cccccc}
954: 1 & y_1 & y_2 & y_3 & y_4 & y_5 \\
955: y_1 & y_3 & y_4 & y_6 & 1 & y_7 \\
956: y_2 & y_4 & y_5 & 1 & y_7 & y_8 \\
957: y_3 & y_6 & 1 & y_9 & y_1 & y_2 \\
958: y_4 & 1 & y_7 & y_1 & y_2 & y_{10}\\
959: y_5 & y_7 & y_8 & y_2 & y_{10} & y_{11}
960: \end{array} \right)
961: \end{array}
962: $$
963: If $M_{\B_2}(\yy) \succeq 0$, then the principal minor indexed by
964: $x_1$ and $x_1x_2$ implies that $y_2y_3 \geq 1$ and so in particular,
965: $y_2 \neq 0$ for all $\yy \in Q_{2}(I)$. However, since $Q_{2}(I)
966: \supseteq \textup{conv}(\V_{\RR}(I)) = \{(s_1,s_2) \in \RR^2 \,:\, s_2
967: > 0 \}$, it must be that $Q_{2}(I) = \textup{conv}(\V_{\RR}(I))$ which
968: shows that $Q_{2}(I)$ is not closed.
969: \end{example}
970:
971: \begin{remark}\label{rem:running example}
972: Example \ref{ex:runningex2} can be modified to show that $Q_k(I)$
973: may not be closed even if $\V_{\RR}(I)$ is finite. To see this,
974: choose sufficiently many pairs of points $(\pm t,1/t^2)$ on the
975: curve $x_1^2x_2=1$ to form a set $S$ such that the ideal $\I(S)$ has a
976: monomial basis $\B'$ in which $\B'_4$ equals the $\B_4$ from
977: above. For instance, $S = \{(\pm t, 1/t^2) \,:\, t=1,\ldots,7 \}$
978: will work. Then $Q_{2}(\I(S))$ coincides with $Q_{2}(I)$ computed
979: above and so is not a closed set.
980: \end{remark}
981:
982:
983: We now show that in the particular case of vanishing ideals of $0/1$
984: points, which are real radical ideals, the closure in
985: Theorem~\ref{thm:basis free formulation} ($\textup{TH}_k(I) =
986: \textup{cl}(Q_k(I))$) is not needed. Most ideals that occur in
987: combinatorial optimization have this form and we will see important
988: examples in Section~\ref{sec:examples}. Remark \ref{rem:running
989: example} shows that the closure cannot be removed for arbitrary
990: finite point sets.
991:
992: \begin{proposition} \label{prop:theta=Q} If $S$ is a set of $0/1$
993: points in $\RR^n$ and $I=\I(S)$ then for all positive integers $k$,
994: $\textup{TH}_k(I)=Q_{k}(I)$.
995: \end{proposition}
996:
997: \begin{proof}
998: By Corollary~\ref{cor:theta-lasserre} it is enough to show that
999: there is a linear polynomial $g \in \RR[\xx]$ such that $g \equiv
1000: \ff_k^t A \ff_k$ mod $I$ for a {\em positive definite} matrix $A$
1001: and some basis of $\RR[\xx]/I$ with respect to which $\ff_k$ was
1002: determined. Let $\B$ be a monomial basis for $\RR[\xx]/I$ and
1003: $\B_k=\{1,p_1,\ldots,p_l\}+I$. Let $\cc \in \RR^{l}$ be the vector
1004: with all entries equal to $-2$, and $D \in \RR^{l \times l}$ be the
1005: diagonal matrix with all diagonal entries equal to $4$. Since $x_i^2
1006: \equiv x_i$ mod $I$ for $i=1,\ldots,n$ and $\B$ is a monomial basis,
1007: for any $f+I \in \B$, $f \equiv f^2$ mod $I$. Therefore, the
1008: constant
1009: $$l+1 \equiv \ff_k^t
1010: \left[
1011: \begin{array} {cc}
1012: l+1 & \cc^t \\
1013: \cc & D
1014: \end{array}
1015: \right]\ff_k \,\,\textup{mod} \,\,I,$$ and it is enough to prove that
1016: the square matrix on the right is positive definite. This follows from
1017: the fact that $D$ is positive definite and its {\em Schur
1018: complement} $(l+1) - \cc^t D^{-1} \cc = 1$ is positive
1019: (\cite[Theorem 7.7.6]{HornJohnson}).
1020: \end{proof}
1021:
1022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1023: \section{Combinatorial Examples} \label{sec:examples}
1024:
1025: An important area of application for the theta body hierarchy
1026: constructed in Section~\ref{sec:theta bodies} is {\em combinatorial
1027: optimization} which is typically concerned with optimizing a linear
1028: function over a finite set of integer points. In this section, we
1029: compute theta bodies for two important problems in combinatorial
1030: optimization -- the {\em maximum stable set problem} and the {\em
1031: maximum cut problem} in a graph. We explain the observations about
1032: the stable set problem which motivated Lov{\'a}sz to pose
1033: Problem~\ref{prob:lovasz}. The cut problem is modeled in two different
1034: ways. The first is a non-standard approach which is described
1035: fully. For the second, more standard model of the cut problem, theta
1036: bodies provide a new hierarchy of semidefinite relaxations for the
1037: {\em cut polytope} that is studied in detail in \cite{GLPT}. We
1038: outline those results briefly here. A recent trend in theoretical
1039: computer science has been to study the computational complexity of
1040: approximating problems in combinatorial optimization via the standard
1041: hierarchies of convex relaxations to these problems such as those in
1042: \cite{LovaszSchrijver91} and \cite{Lasserre1, Lasserre2}. Our theta
1043: body approach provides a new mechanism to establish such complexity
1044: results.
1045:
1046: \subsection{The Maximum Stable Set Problem}\label{subsec:stable sets}
1047: Let $G=([n],E)$ be an undirected graph with vertex set
1048: $[n]=\{1,\ldots,n\}$ and edge set $E$. A {\em stable set} in $G$ is a
1049: set $U \subseteq [n]$ such that for all $i,j \in U$, $\{i,j\} \not \in
1050: E$. The maximum stable set problem seeks the stable set of largest
1051: cardinality in $G$, the size of which is the {\em stability number} of
1052: $G$, denoted as $\alpha(G)$.
1053:
1054: The maximum stable set problem can be modeled as follows. For each
1055: stable set $U \subseteq [n]$, let $\chi^U \in \{0,1\}^n$ be its {\em
1056: characteristic vector} defined as $(\chi^U)_i = 1$ if $i \in U$ and
1057: $(\chi^U)_i = 0$ otherwise. Let $S_G \subseteq \{0,1\}^n$ be the set
1058: of characteristic vectors of all stable sets in $G$. Then
1059: $\textup{STAB}(G) := \conv(S_G)$ is called the {\em stable set
1060: polytope} of $G$ and the maximum stable set problem is, in theory,
1061: the linear program $\textup{max}\{ \sum_{i=1}^{n} x_i \,:\, \xx \in
1062: \textup{STAB}(G) \}$ with optimal value $\alpha(G)$. However,
1063: $\textup{STAB}(G)$ is not known apriori, and so one resorts to
1064: relaxations of it over which one can optimize $\sum_{i=1}^{n} x_i$.
1065:
1066: In \cite{ShannonCapacity}, Lov{\'a}sz introduced, $\textup{TH}(G)$, a
1067: convex relaxation of $\textup{STAB}(G)$, called the {\em theta body}
1068: of $G$. The problem $\textup{max}\{ \sum_{i=1}^{n} x_i \,:\, \xx \in
1069: \textup{TH}(G) \}$ is a SDP which can be solved to arbitrary precision
1070: in polynomial time in the size of $G$. The optimal value of this SDP
1071: is called the {\em theta number} of $G$ and provides an upper
1072: bound on $\alpha(G)$. See \cite[Chapter 9]{GLS} and
1073: \cite{BruceShepherd} for more on the stable set problem and
1074: $\textup{TH}(G)$. The body $\textup{TH}(G)$ was the first example of
1075: a SDP relaxation of a discrete optimization problem and snowballed the
1076: use of SDP in combinatorial optimization. See \cite{LaurentRendl,
1077: Lovasz} for surveys. Recall that a graph $G$ is {\em perfect} if and
1078: only if $G$ has no induced odd cycles of length at least five or their
1079: complements. Lov{\'a}sz showed that $\textup{STAB}(G) =
1080: \textup{TH}(G)$ if and only if $G$ is perfect. This equality shows
1081: that the maximum stable set problem can be solved in polynomial time
1082: in the size of $G$ when $G$ is a perfect graph, and this geometric
1083: proof is the only one known for this complexity result.
1084:
1085: The theta body $\textup{TH}(G)$ has many definitions (see
1086: \cite[Chapter 9]{GLS}) but the one relevant for this paper was
1087: observed by Lov{\'a}sz and appears without proof in
1088: \cite{LovaszStablesetsPolynomials}. Let $I_G := \langle x_j^2 - x_j
1089: \,\,\forall\,\,j \in [n], \,\,\, x_ix_j \,\, \forall \,\, \{i,j\} \in
1090: E \rangle \subseteq \RR[\xx]$. Then check that $\V_{\RR}(I_G) = S_G$
1091: and that $I_G$ is both zero-dimensional and real radical. Lov{\'a}sz
1092: observed that
1093: \begin{equation} \label{lovasz observation}
1094: \textup{TH}(G) = \{ \xx \in \RR^n \,:\, f(\xx) \geq 0 \,\,\forall
1095: \,\,\textup{ linear } \,\,f\,\,\textup{that is $1$-sos mod} \,\,I_G
1096: \}.
1097: \end{equation}
1098: By Definition~\ref{def:theta} (1), $\textup{TH}(G)$ is exactly the
1099: first theta body, $\textup{TH}_1(I_G)$, of the ideal $I_G$, and by the
1100: above discussion, $I_G$ is $\textup{TH}_1$-exact (i.e.,
1101: $\textup{TH}_1(I_G) = \textup{STAB}(G)$) if and only if $G$ is
1102: perfect. Lov{\'a}sz observed that, in fact, $I_G$ is $(1,1)$-sos if
1103: and only if $G$ is perfect which motivated Problem~\ref{prob:lovasz}
1104: that asks for a characterizations of all $(1,1)$-sos ideals in
1105: $\RR[\xx]$. Lov{\'a}sz refers to a $(1,1)$-sos ideal as a {\em
1106: perfect ideal}. A $(1,1)$-sos ideal $I$ would have the property that
1107: its first and simplest theta body, $\textup{TH}_1(I)$, coincides with
1108: $\textup{cl}(\conv(\V_\RR(I)))$ which is a valuable property for
1109: linear optimization over $\conv(V_\RR(I))$, especially when
1110: $\textup{TH}_1(I)$ is computationally tractable.
1111:
1112: The theta body hierarchy of the ideal $I_G$ therefore naturally
1113: extends the theta body of $G$ to a family of nested relaxations of
1114: $\textup{STAB}(G)$. Further, the connection between $\textup{TH}(G)$
1115: and sums of squares polynomials motivated Definition~\ref{def:theta}
1116: which extends the construction of $\textup{TH}(G)$ to a hierarchy of
1117: relaxations of $\V_\RR(I)$ for any ideal $I \subseteq \RR[\xx]$. We
1118: now explicity describe the $k$-th theta body of $I_G$ in terms of
1119: combinatorial moment matrices.
1120:
1121: For $U \subseteq [n]$, let $\xx^U := \prod_{i \in U} x_i$. From the
1122: generators of $I_G$ it is clear that if $f \in \RR[\xx]$, then $f
1123: \equiv g$ mod $I_G$ where $g$ is in the $\RR$-span of the set of
1124: monomials $\{ \xx^U \,:\, U \textup{ is a stable set in } G \}$. Check
1125: that $\B := \{\xx^U + I_G \,:\, U \,\,\textup{stable set in} \,\, G\}$
1126: is a basis of $\RR[\xx]/I_G$ containing $1+I_G, x_1 + I_G, \ldots, x_n
1127: + I_G$. Therefore, by Corollary~\ref{cor:simple version} and
1128: Proposition~\ref{prop:theta=Q} we have
1129: $$\textup{TH}_k(I_G) = \left\{ \yy \in \RR^n \,:\,
1130: \begin{array}{l}
1131: \exists \, M \succeq 0, \, M \in \RR^{|\B_k| \times |\B_k|}
1132: \,\textup{such that} \\
1133: M_{\emptyset \emptyset} = 1,\\
1134: M_{\emptyset \{i\}} = M_{\{i\} \emptyset} = M_{\{i\} \{i\}} = y_i \\
1135: M_{U U'} = 0 \,\,\textup{if} \,\,U \cup U' \,\, \textup{is not stable
1136: in} \,\, G\\
1137: M_{U U'} = M_{W W'} \,\,\textup{if} \,\, U \cup U' = W \cup W'
1138: \end{array}
1139: \right \}.$$
1140: In particular, indexing the one element stable sets by the vertices
1141: of $G$,
1142: $$\textup{TH}_1(I_G) = \left\{ \yy \in \RR^n \,:\,
1143: \begin{array}{l}
1144: \exists \, M \succeq 0, M \in \RR^{(n+1) \times (n+1)}\,\textup{such that} \\
1145: M_{00} = 1,\\
1146: M_{0i} = M_{i0} = M_{ii} = y_i \,\,\forall\,\,i \in [n]\\
1147: M_{ij} = 0 \,\,\forall \,\, \{i,j\} \in E
1148: \end{array}
1149: \right \}.$$
1150:
1151: This description of $\textup{TH}_1(I_G)$ coincides with the
1152: semidefinite description of $\textup{TH}(G)$ (see \cite[Lemma
1153: 2.17]{LovaszSchrijver91} for instance) and so, $\textup{TH}(G) =
1154: \textup{TH}_1(I_G)$. Corollary~\ref{cor:real radical equivalence}
1155: confirms Lov{\'a}sz's observation and adds to his other
1156: characterizations of a perfect graph as follows.
1157:
1158: \begin{theorem} \cite[Chapter 9]{GLS} \label{thm:knownforperfectgraphs}
1159: The following are equivalent for a graph $G$.
1160: \begin{enumerate}
1161: \item $G$ is perfect.
1162: \item $\textup{STAB}(G) = \textup{TH}(G)$.
1163: \item $\textup{TH}(G)$ is a polytope.
1164: \item The complement $\overline G$ of $G$ is perfect.
1165: \item $I_G$ is $(1,1)$-sos.
1166: \end{enumerate}
1167: \end{theorem}
1168:
1169: The usual Lasserre relaxations of the maximum stable set problem are
1170: set up from the following initial linear programming relaxation of
1171: $\textup{STAB}(G)$:
1172: $$\textup{FRAC}(G) := \{ \xx \in \RR^n \,:\, x_i \geq 0 \,\,\forall
1173: \,\,i \in [n], \, 1-x_i-x_j \geq 0 \,\,\forall \,\,\{i,j\} \in E \}.$$
1174: Note that $S_G = \textup{FRAC}(G) \cap \{0,1\}^n$. The $k$-th
1175: Lasserre relaxation of $\textup{STAB}(G)$ (see \cite{Lasserre2},
1176: \cite{LaurentComparisonpaper}) uses both the ideal $\langle x_i^2-x_i
1177: \,:\, i \in [n] \rangle$ and the inequality system describing
1178: $\textup{FRAC}(G)$, whereas in the theta body formulation,
1179: $\textup{TH}_k(I_G)$, there is only the ideal $I_G$ and no
1180: inequalities. Despite this difference, \cite[Lemma
1181: 20]{LaurentComparisonpaper} proves that the usual Lasserre hierarchy
1182: is exactly our theta body hierarchy for the stable set problem. This
1183: interpretation of the Lasserre hierarchy provides new tools to
1184: understand these relaxations such as establishing the validity of
1185: inequalities over them as shown below.
1186:
1187: Since no monomial in the basis $\B$ of $\RR[\xx]/I_G$ has degree
1188: larger than $\alpha(G)$, for any $G$, $I_G$ is $(1,\alpha(G))$-sos and
1189: $\textup{STAB}(G) = \textup{TH}_{\alpha(G)}(I_G)$. However, for many
1190: non-perfect graphs the theta-rank of $I_G$ can be a lot smaller than
1191: $\alpha(G)$. For instance if $G$ is a $(2k+1)$-cycle, then $\alpha(G)
1192: = k$ while Proposition~\ref{prop:2sos} below shows that the theta-rank
1193: of $I_G$ is two.
1194:
1195:
1196: \begin{theorem} \label{thm:stabg for oddholes}
1197: \cite[Corollary 65.12a]{SchrijverB} If $G=([n],E)$ is an odd cycle
1198: with $n \geq 5$, then $\textup{STAB}(G)$ is determined by the
1199: following inequalities:
1200: $$
1201: x_i \geq 0 \,\,\forall \,\, i \in [n], \,\,\, 1 - \sum_{i \in K}
1202: x_i \geq 0 \,\,\forall \textup{ cliques } \,\,K \textup{ in } G,
1203: \,\,\, \alpha(G) - \sum_{i \in [n]} x_i \geq 0.
1204: $$
1205: \end{theorem}
1206:
1207: \begin{proposition} \label{prop:2sos}
1208: If $G$ is an odd cycle with at least five vertices, then $I_G$ is
1209: $(1,2)$-sos and therefore, $\textup{TH}_2$-exact.
1210: \end{proposition}
1211:
1212: \begin{proof}
1213: Let $n = 2k+1$ and $G$ be an $n$-cycle. Then $I_G = \langle
1214: x_i^2-x_i, \,\, x_ix_{i+1} \,\,\forall \,\, i \in [n] \rangle$ where
1215: $x_{n+1} = x_1$. Therefore, $(1-x_i)^2 \equiv 1-x_i$ and
1216: $(1-x_i-x_{i+1})^2 \equiv 1 -x_i-x_{i+1} \,\,\textup{mod}\,\,I_G$.
1217: This implies that, mod $I_G$,
1218: $$p_i^2:=((1-x_1)(1-x_{2i}-x_{2i+1}))^2 \equiv p_i = 1 - x_1
1219: -x_{2i}-x_{2i+1} + x_1x_{2i} + x_1x_{2i+1}.$$
1220: Summing over $i=1,..,k$, we get
1221: $$\sum_{i=1}^{k} p_i^2 \equiv k - kx_1 -\sum_{i=2}^{2k+1}x_i +
1222: \sum_{i=3}^{2k} x_1x_i \,\,\textup{mod}\,\,I_G$$
1223: since $x_1x_2$ and $x_1x_{2k+1}$ lie in
1224: $I_G$. Define $g_i := x_1(1-x_{2i+1}-x_{2i+2})$. Then $g_i^2 - g_i
1225: \in I_G$ and mod $I_G$ we get that
1226: $$\sum_{i=1}^{k-1} g_i^2 \equiv (k-1)x_1 - \sum_{i=3}^{2k} x_1x_i,
1227: \,\,\textup{which implies} \,\, \sum_{i=1}^{k} p_i^2 +
1228: \sum_{i=1}^{k-1} g_i^2 \equiv k - \sum_{i=1}^{2k+1}x_i.$$
1229:
1230: To prove that $I_G$ is $(1,2)$-sos it suffices to show that the left
1231: hand sides of the inequalities in the description of
1232: $\textup{STAB}(G)$ in Theorem~\ref{thm:stabg for oddholes} are $2$-sos
1233: mod $I_G$ since by Farkas Lemma \cite{Sch}, all other linear
1234: inequalities that are non-negative over $S_G$ are non-negative real
1235: combinations of a set of inequalities defining $\textup{STAB}(G)$.
1236: Clearly, $x_i \equiv x_i^2 \,\,\textup{mod}\,\,I_G$ for all $i \in
1237: [n]$ and one can check that for each clique $K$, $(1 - \sum_{i \in K}
1238: x_i) \equiv (1 - \sum_{i \in K} x_i)^2 \,\,\textup{mod}\,\,I_G$. The
1239: previous paragraph shows that $k - \sum_{i=1}^{2k+1}x_i$ is also
1240: $2$-sos mod $I_G$.
1241: \end{proof}
1242:
1243: An induced odd cycle $C_{2k+1}$ in $G$, yields the well-known {\em odd
1244: cycle inequality} $\sum_{i\in C_{2k+1}} x_i \leq \alpha(C_{2k+1})=k$
1245: that is satisfied by $S_G$ \cite[Chapter
1246: 9]{GLS}. Proposition~\ref{prop:2sos} implies that for any graph $G$,
1247: $\textup{TH}_2(I_G)$ satisfies all odd cycle inequalities from $G$
1248: since every stable set $U$ in $G$ restricts to a stable set in an
1249: induced odd cycle in $G$. This general result can also be proved using
1250: results from \cite{LovaszSchrijver91} and
1251: \cite{LaurentComparisonpaper}. The direct arguments used in the proof
1252: of Proposition~\ref{prop:2sos} are examples of the algebraic {\em
1253: inference rules} outlined by Lov{\'a}sz in
1254: \cite{LovaszStablesetsPolynomials}. Similarly, one can also show that
1255: other well-known classes of inequalities such as the {\em odd
1256: antihole} and {\em odd wheel} inequalities \cite[Chapter 9]{GLS} are
1257: also valid for $\textup{TH}_2(I_G)$. Schoenebeck \cite{Schoenebeck}
1258: has recently shown that there is no constant $k$ such that
1259: $\textup{STAB}(G) = \textup{TH}_k(I_G)$ for all graphs $G$ (as
1260: expected, unless P=NP). However, no explicit family of graphs that
1261: exhibit this behaviour is known.
1262:
1263: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1264: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1265:
1266: \subsection{Cuts in graphs} \label{subsec:cuts} Given an undirected
1267: connected graph $G = ([n], E)$ and a partition of its vertex set $[n]$
1268: into two parts $V_1$ and $V_2$, the set of edges $\{i,j\} \in E$ such
1269: that exactly one of $i$ or $j$ is in $V_1$ and the other in $V_2$ is
1270: the {\em cut} in $G$ induced by the partition $(V_1,V_2)$. The cuts in
1271: $G$ are in bijection with the $2^{n-1}$ distinct partitions of $[n]$
1272: into two sets. The maximum cut problem in $G$ seeks the cut in $G$ of
1273: largest cardinality. This problem is NP-hard and has received a great
1274: deal of attention in the literature. A celebrated result in this area
1275: is an approximation algorithm for the max cut problem, due to Goemans
1276: and Williamson \cite{GoemansWilliamson}, that guarantees a cut of size
1277: at least $0.878$ of the optimal cut. It relies on a simple SDP
1278: relaxation of the problem.
1279:
1280: We first study a non-standard model of the max cut problem. Let
1281: $$SG := \{ \chi^F \,:\, F \subseteq E\,\,\textup{is contained in a cut
1282: of} \,\, G \} \subseteq \{0,1\}^E.$$ Then the {\em weighted} max cut
1283: problem with non-negative weights $w_e$ on the edges $e \in E$ is
1284: $\textup{max} \left \{ \sum_{e \in E} w_e x_e \,:\, \xx \in SG \right
1285: \}$, and the vanishing ideal
1286: $$\I(SG) = \langle x_e^2 - x_e \,\,, \xx^T \,:\,
1287: \,\,e \in E, \,\, T \textup{ odd cycle in } G \rangle. $$ A
1288: basis of $\RR[\xx]/{\I(SG)}$ is $$\B = \{ \xx^U + I(SG) \,:\, U
1289: \subseteq E \,\,\textup{does not contain an odd cycle in} \,\, G \}$$
1290: and $1+\I(SG), x_e + \I(SG) \,(\forall \,e \in E)$ lie in
1291: $\B$. Therefore,
1292: $$\textup{TH}_k(\I(SG)) = \left\{ \yy \in \RR^E \,:\,
1293: \begin{array}{l}
1294: \exists \, M \succeq 0, \, M \in \RR^{|\B_k| \times |\B_k|}
1295: \,\textup{such that} \\
1296: M_{\emptyset \emptyset} = 1,\\
1297: M_{\emptyset \{i\}} = M_{\{i\} \emptyset} = M_{\{i\}\{i\}} = y_i \\
1298: M_{U U'} = 0 \,\,\textup{if} \,\,U \cup U' \,\, \textup{has an odd
1299: cycle} \\
1300: M_{U U'} = M_{W W'} \,\,\textup{if} \,\, U \cup U' = W \cup W'
1301: \end{array}
1302: \right \}.$$
1303:
1304: In particular,
1305: $$\textup{TH}_1(\I(SG)) = \left\{ \yy \in \RR^E \,:\,
1306: \begin{array}{l}
1307: \exists \, M \succeq 0, \, M \in \RR^{(|E|+1) \times (|E|+1)}
1308: \,\textup{such that} \\
1309: M_{00} = 1,\\
1310: M_{0 e} = M_{e 0} = M_{e e} = y_e \,\,\forall \,\, e \in E
1311: \end{array}
1312: \right \}.$$
1313:
1314: Note that for any graph $G$, $\textup{TH}_1(\I(SG))$ is the unit cube
1315: in $\RR^{E}$ which may not be equal to $\textup{conv}(SG)$. This
1316: stands in contrast to the case of stable sets for which
1317: $\textup{TH}_1(I_G)$ is a polytope if and only if $\textup{TH}_1(I_G)
1318: = \textup{STAB}(G)$.
1319:
1320: \begin{proposition}
1321: The ideal $\I(SG)$ is $\textup{TH}_1$-exact if and only if $G$ is a
1322: bipartite graph.
1323: \end{proposition}
1324:
1325: \begin{proof} This follows immediately from the description of
1326: $\textup{TH}_1(\I(SG))$ and from the fact that $G$ is bipartite if
1327: and only if it has no odd cycles.
1328: \end{proof}
1329:
1330: Since the maximum degree of a monomial in $\B$ is the size of the max
1331: cut in $G$, the theta-rank of $\I(SG)$ is bounded from above by the
1332: size of the max cut in $G$.
1333:
1334: \begin{proposition} \label{prop:increasingk} There is no constant $k$
1335: such that $\I(SG)$ is $\textup{TH}_k$-exact for all graphs $G$.
1336: \end{proposition}
1337:
1338: \begin{proof}
1339: Let $G$ be a $(2k+1)$-cycle. Then $\textup{TH}_k(\I(SG)) \neq
1340: \textup{conv}(SG)$ since the linear constraint imposed by the cycle
1341: in the definition of $\textup{TH}_k(\I(SG))$ will not appear in
1342: theta bodies of index $k$ or less.
1343: \end{proof}
1344:
1345: % If $G$ is a $(2m+1)$-cycle then in fact,
1346: % $\textup{TH}_1(\I(SG)) = \textup{TH}_2(\I(SG)) = \cdots =
1347: % \textup{TH}_m(\I(SG))$ is the unit cube in $\RR^E$ even though
1348: % $\textup{TH}_m(\I(SG)) \neq \textup{conv}(SG)$. This example shows
1349: % that the $k$-th theta body need not strictly contain the $(k+1)$-th
1350: % body in the theta body hierarchy.
1351:
1352: The theta bodies of a second, more standard, formulation of the
1353: weighted max cut problem are studied in \cite{GLPT}. In this setup,
1354: each cut $C$ in $G=([n],E)$ is recorded by its {\em cut vector}
1355: $\chi^C \in \{\pm 1\}^{E}$ with $\chi^C_{\{i,j\}} = 1$ if $\{i,j\}
1356: \not \in C$ and $\chi^C_{\{i,j\}} = -1$ if $\{i,j\} \in C$. Let $E_n$
1357: denote the edge set of the complete graph $K_n$, and $\pi_E$ be the
1358: projection from $\RR^{E_n}$ to $\RR^E$. The {\bf cut polytope} of $G$
1359: is
1360: $$\textup{CUT}(G) := \textup{conv}\{ \chi^C : C \textrm{ is a cut in }
1361: G\} \subseteq \RR^{E} = \pi_E(\textup{CUT}(K_n)),$$
1362: and the weighted
1363: max cut problem, for weights $w_e \in \RR$ ($\forall$ $e \in E$) becomes
1364: $$
1365: \max \left\{\frac{1}{2}\sum_{e \in E} w_{e}(1- x_{e}) : \xx \in
1366: \textup{CUT}(G)\right\}.$$
1367:
1368: In \cite{GLPT}, the vanishing ideal $IG$ of the cut vectors $\{ \chi^C
1369: \,:\, C \textup{ is a cut in } G \}$ is described and a combinatorial
1370: basis $\B$ for $\RR E/ IG$ is identified. Using these, the $k$-th
1371: theta body, $\textup{TH}_k(IG)$, of $IG$ can be described as:
1372: $$\left\{ \yy \in \RR^E \,:\,
1373: \begin{array}{l}
1374: \exists \, M \succeq 0, \, M \in \RR^{|\B_k| \times
1375: |\B_k|}\,\,\textup{such that} \\
1376: M_{\emptyset,\emptyset} = 1\\
1377: M_{F_1,F_2} = M_{F_3,F_4} \,\textup{ if } F_1 \Delta F_2 \Delta F_3
1378: \Delta F_4 \,\textup{ is a cycle in } G
1379: \end{array}
1380: \right \}.$$
1381: % \begin{definition} \label{def:T-join}
1382: % Let $G=([n],E)$ be an undirected graph, $T \subseteq [n]$ have even
1383: % cardinality and $F \subseteq E$. Then $F$ is called a {\bf $T$-join}
1384: % if $T$ is exactly the set of vertices of odd-degree in the subgraph
1385: % of $G$ induced by $F$.
1386: % \end{definition}
1387:
1388: % Let $\RR E := \RR[x_e \,:\, e \in E]$ and $IG \subseteq \RR E$ be the
1389: % vanishing ideal of the cut vectors in $G$. Identify a subgraph $F$ of
1390: % $G$ with its set of edges $E(F) \subseteq E$. For a subgraph $F$ in
1391: % $G$, let $\xx^F$ denote the squarefree monomial
1392: % $$\prod_{e \in E(F)} x_e \in \RR E.$$
1393:
1394: % \begin{theorem} \cite[Theorem 3.4]{GLPT} \label{thm:basis of RG/IG}
1395: % The ideal $IG$ is generated by the binomials $x_e^2-1$, $\forall$ $e
1396: % \in E$ and $1-\xx^D$, $\forall$ chordless circuits $D$ in $G$. For
1397: % an even subset $T \subseteq [n]$, let $F_T$ be a $T$-join in $G$ and
1398: % let $\B := \{ \xx^{F_T} + IG \,:\, T \subseteq [n], \,|T|
1399: % \,\textup{even} \}$. Then $\B$ is a basis of $\RR E/IG$.
1400: % \end{theorem}
1401:
1402: % Set $\B$ as in Theorem~\ref{thm:basis of RG/IG} with the additional
1403: % stipulation that for each $T \subseteq [n]$ of even cardinality, the
1404: % $T$-join $F_T$ chosen to represent a coset in $\B$ is one with
1405: % smallest number of edges among all $T$-joins in $G$. Index a coset
1406: % $\xx^{F}+IG \in \B$ by $F$ and let $F_1 \Delta F_2$ denote the
1407: % symmetric difference of the sets $F_1$ and $F_2$. For each positive
1408: % integer $k$, let $\B_k$ be the elements of $\B$ of degree at most $k$
1409: % which are exactly the cosets in $\B$ given by $T$-joins of size at
1410: % most $k$. When $G$ is a simple graph (i.e., $G$ has no loops or
1411: % parallel edges) we show in Section~3 of \cite{GLPT} that the
1412:
1413: % and in particular,
1414:
1415: % $$\textup{TH}_1(IG) = \left\{ \yy \in \RR^E \,:\,
1416: % \begin{array}{l}
1417: % \exists \, M \succeq 0, \, M \in \RR^{(|E|+1) \times
1418: % (|E|+1)}\,\,\textup{such that} \\
1419: % M_{\emptyset,\emptyset} = M_{e,e} = 1 \,\,\forall \,\,e \in E \\
1420: % M_{e,f} = M_{\emptyset,g} \,\,\textup{ if } \,\,\{e,f,g\} \,\,\textup{
1421: % is a triangle in } \,\,G, \\
1422: % M_{e,f} = M_{g,h} \,\, \textup{ if }\,\,\{e,f,g,h\} \,\,\textup{ is a
1423: % circuit in }\,\,G
1424: % \end{array}
1425: % \right \}.$$
1426:
1427: % In Section~3.3 of \cite{GLPT} we compare the theta bodies shown above
1428: % to other known relaxations of $\textup{CUT}(G)$ such as those in
1429: % \cite{LaurentMaxCut}. In particular, we show that if $G^*$ is the
1430: % suspension of $G$ from a new vertex then $\pi_E(\textup{TH}_1(IG^*))$
1431: % is contained in the Goemans-Williamson relaxation of
1432: % $\textup{CUT}(G)$.
1433: These theta bodies provide a new canonical set of SDP relaxations for
1434: $\textup{CUT}(G)$ that exploits the structure of $G$ directly. It is
1435: also shown in \cite{GLPT} that $IG$ is $\textup{TH}_1$-exact if and
1436: only if $G$ has no $K_5$-minor and no induced cycle of length at
1437: least five which answers Problem~8.4 posed by Lov{\'a}sz in
1438: \cite{Lovasz}.
1439:
1440: \begin{remark}
1441: We remark that the stable set problem and the first formulation of
1442: the max cut problem discussed above are special cases of the
1443: following general setup. Let $\Delta$ be an {\em abstract
1444: simplicial complex} (or {\em independence system}) with vertex set
1445: $[n]$ recorded as a collection of subsets of $[n]$, called the {\em
1446: faces} of $\Delta$. The {\em Stanley-Reisner} ideal of $\Delta$
1447: is the ideal $J_\Delta$ generated by the squarefree monomials
1448: $x_{i_1}x_{i_2} \cdots x_{i_k}$ such that $\{i_1, i_2, \ldots, i_k\}
1449: \subseteq [n]$ is not a face of $\Delta$. If $I_\Delta := J_\Delta
1450: + \langle x_i^2 - x_i \,:\, i \in [n] \rangle$, then
1451: $\V_{\RR}(I_\Delta) = \{ \ss \in \{0,1\}^n \,:\,
1452: \textup{support}(\ss) \in \Delta \}$. For $T \subseteq [n]$, recall
1453: that $\xx^T := \prod_{i \in T} x_i$. Then $\B := \{ \xx^T \,:\, T
1454: \in \Delta \} + I_{\Delta}$ is a basis for $\RR[\xx]/I_\Delta$
1455: containing $1+I_\Delta, x_1+I_\Delta, \ldots, x_n+I_\Delta$.
1456: Therefore, by Corollary~\ref{cor:simple version} and
1457: Proposition~\ref{prop:theta=Q}, the $k$-th theta body of $I_\Delta$
1458: is
1459: $$\textup{TH}_k(I_\Delta) = \textup{proj}_{y_1, \ldots, y_n} \{ \yy
1460: \in \RR^{\B_{2k}} \,:\, M_{\B_k}(\yy) \succeq 0, \, y_0 = 1\}.$$
1461: Since
1462: $\B$ is in bijection with the faces of $\Delta$, and $x_i^2 - x_i \in
1463: I_\Delta$ for all $i \in [n]$, the theta body can be written
1464: explicitly as follows:
1465: $$\textup{TH}_k(I_\Delta) = \left\{ \yy \in \RR^n \,:\,
1466: \begin{array}{l}
1467: \exists \, M \succeq 0, \, M \in \RR^{|\B_k| \times |\B_k|}
1468: \,\textup{such that} \\
1469: M_{\emptyset\emptyset} = 1,\\
1470: M_{\emptyset \{i\}} = M_{\{i\} \emptyset} = M_{\{i\}\{i\}} = y_i \\
1471: M_{UU'} = 0 \,\,\textup{if} \,\,U \cup U' \not \in \Delta\\
1472: M_{UU'} = M_{W W'} \,\,\textup{if} \,\, U \cup U' = W \cup W'
1473: \end{array}
1474: \right \}.$$
1475: If the dimension of $\Delta$ is $d-1$ (i.e., the largest faces in
1476: $\Delta$ have size $d$), then $I_\Delta$ is $(1,d)$-sos and therefore,
1477: $\textup{TH}_d$-exact since all elements of $\B$ have degree at most
1478: $d$. However, the theta-rank of $I_\Delta$ could be much less than
1479: $d$.
1480: \end{remark}
1481:
1482: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1483:
1484: \section{Vanishing ideals of finite
1485: sets of points} \label{sec:structure}
1486:
1487:
1488: Recall that when $S \subset \RR^n$ is finite, its vanishing ideal
1489: $\I(S)$ is zero-dimensional and real radical.
1490:
1491: \begin{definition} \label{def:exact} We say that a finite set
1492: $S \subset \RR^n$ is {\it exact} if its vanishing ideal $\I(S)
1493: \subseteq \RR[\xx]$ is $\textup{TH}_1$-exact.
1494: \end{definition}
1495:
1496: We now answer Lov{\'a}sz's question (Problem~\ref{prob:lovasz}) for
1497: vanishing ideals of finite point sets in $\RR^n$.
1498:
1499: \begin{theorem} \label{thm:perfect}
1500: For a finite set $S \subset \RR^n$, the following are equivalent.
1501: \begin{enumerate}
1502: \item $S$ is exact.
1503: \item $\I(S)$ is $(1,1)$-sos.
1504: \item There is a linear inequality description of $\textup{conv}(S)$,
1505: of the form $g_i(x) \geq 0 \ (i=1,\ldots,m),$ where each $g_i$ is
1506: $1$-sos mod $\I(S)$.
1507: \item There is a linear inequality description of $\textup{conv}(S)$,
1508: of the form $g_i(x) \geq 0 \ (i=1,\ldots,m),$ where each $g_i$ is an
1509: idempotent mod $\I(S)$, i.e., $g_i^2-g_i \in \I(S)$ for
1510: $i=1,\ldots,m$.
1511: \item There is a linear inequality description of $\textup{conv}(S)$,
1512: of the form $g_i(x) \geq 0 \ (i=1,\ldots,m),$ where each $g_i$ takes
1513: at most two different values in $S$, i.e., for each $i$, $S$ is
1514: contained in the union of the hyperplane $g_i(\xx)=0$ and one unique
1515: parallel translate of it.
1516: \end{enumerate}
1517: \end{theorem}
1518:
1519: \begin{proof}
1520: Since $\I(S)$ is real radical, by Corollary~\ref{cor:real radical
1521: equivalence}, (1) $\Leftrightarrow$ (2).
1522:
1523: The implication (2) $\Rightarrow$ (3) follows from the fact that
1524: $\textup{conv}(S)$ has a finite linear inequality description, since
1525: $S$ is finite. The implication (3) $\Rightarrow$ (2) follows from
1526: Farkas lemma, which implies that any valid inequality on $S$ is a
1527: non-negative real combination of the linear inequalities $g_i(x)
1528: \geq 0$.
1529:
1530: Suppose (3) holds and $\textup{conv}(S)$ is a full-dimensional
1531: polytope. Let $F$ be a facet of $\textup{conv}(S)$, and $g(\xx)\geq
1532: 0$ its defining inequality in the given description of
1533: $\textup{conv}(S)$. Then $g(\xx)$ is $1$-sos mod $\I(S)$ if and
1534: only if there are linear polynomials $h_1,\ldots,h_l \in \RR[\xx]$
1535: such that $g \equiv h_1^2 + \cdots + h_l^2 \mod \I(S)$. In
1536: particular, since $g(\xx)=0$ on the vertices of $F$, and all the
1537: $h_i^2$ are non-negative, each $h_i$ must be zero on all the
1538: vertices of $F$. Hence, since the $h_i$'s are linear, they must
1539: vanish on the affine span of $F$ which is the hyperplane defined by
1540: $g(\xx)=0$. Thus each $h_i$ must be a multiple of $g$ and $g \equiv
1541: \alpha g^2$ mod $\I(S)$ for some $\alpha > 0$. We may assume that
1542: $\alpha = 1$ by replacing $g(\xx)$ by $g'(\xx) := \alpha g(\xx)$.
1543: If $\textup{conv}(S)$ is not full-dimensional, then since mod
1544: $\I(S)$, all linear polynomials can be assumed to define hyperplanes
1545: whose normal vectors are parallel to the affine span of $S$, the
1546: proof still holds. Therefore, (3) implies (4). Conversely, since if
1547: for a linear polynomial $g$, $g \equiv g^2$ mod $\I(S)$, then $g$ is
1548: $1$-sos mod $\I(S)$, (4) implies (3).
1549:
1550: The equivalence (4) $\Leftrightarrow$ (5) follows since $g \equiv
1551: g^2$ mod $\I(S)$ if and only if $g(\ss)(1 - g(\ss)) = 0 \,\,\forall
1552: \,\, \ss \in S$.
1553: \end{proof}
1554:
1555: Recall from the discussion at the end of Section 2.1 that by results
1556: of Parrilo, if $I \subseteq \RR[\xx]$ is a zero-dimensional radical
1557: ideal, then the theta-rank of $I$ is at most $|\V_{\CC}(I)|-1$. Better
1558: upper bounds can be derived using the following extension of Parrilo's
1559: theorem.
1560:
1561: \begin{remark} \label{rem:k parallel translates} Suppose $S \subseteq
1562: \RR^n$ is a finite set such that each facet $F$ of
1563: $\textup{conv}(S)$ has a facet defining inequality $h_F(\xx) \geq 0$
1564: where $h_F$ takes at most $t+1$ values on $S$, then $\I(S)$ is
1565: $\textup{TH}_t$-exact: In this case, it is easy to construct a
1566: degree $t$ intepolator $g$ for the values of $\sqrt{h_F}$ on $S$,
1567: and we have $h_F \equiv g^2$ mod $\I(S)$. The result then follows
1568: from Farkas Lemma.
1569: \end{remark}
1570:
1571:
1572: % \begin{proof}
1573: % Let $H_F$ be defined by the linear equation $\cc_F \cdot \xx =
1574: % a_0^F$ with $\cc_F \cdot \ss \geq a_0^F$ for all $\ss \in S$.
1575: % Suppose $S$ is contained in the hyperplanes with equations $c_F
1576: % \cdot \xx = a_0^F, \, c_F \cdot \xx = a_1^F, \,\ldots,\, c_F \cdot
1577: % \xx = a_{t_F}^F$ and $t_F \leq t$. For a fixed $\ss \in S$, there
1578: % exists $s_j \in \{0,\ldots, t_F\}$ such that $\cc_F \cdot s =
1579: % a_{s_j}^F \geq a_0^F$. For $0 \leq j \leq t_F$, define
1580: % $$
1581: % q_{F,j}(\xx) := \frac{\prod_{i \neq j} (\cc_F \cdot \xx -
1582: % a_i^F)}{\prod_{i \neq j} (a_j^F - a_i^F)} \in \RR[\xx].$$
1583: % Then
1584: % $q_{F,j}(\ss) = 0$ if $j \neq s_j$, $q_{F,s_j}(\ss) = 1$, and
1585: % $q_{F,j}(\xx)$ has degree at most $t$. Suppose $f$ is a linear
1586: % polynomial such that $f \geq 0 \,\,\textup{mod} \,\,\I(S)$. Then
1587: % there exists scalars $\lambda_F \geq 0$ for each facet $F$ of
1588: % $\textup{conv}(S)$ and a linear polynomial $g(\xx) \in \I(S)$ such
1589: % that $$f(\xx) = \sum_F \lambda_F(\cc_F \cdot \xx - a_0^F)+g(\xx).$$
1590: % Define $$h(\xx) := f(\xx) - \sum_F \sum_{j=1}^{t_F}
1591: % \left(\sqrt{\lambda_F (a_j^F - a_0^F)} \,\,q_{F,j}(\xx)\right)^2.$$
1592: % Then for $\ss \in S$, $h(\ss) = f(\ss) - \sum_F \sum_{j=1}^{t_F}
1593: % \left(\sqrt{\lambda_F (a_j^F - a_0^F)} \,\,q_{F,j}(\ss)\right)^2 =
1594: % f(\ss) - \sum_F \lambda_F (a_{s_j}^F - a_0^F) = 0$ which implies
1595: % that $h(\xx) \in \I(S)$ since $\I(S)$ is radical. Therefore, $f$ is
1596: % $t$-sos and thus, $\I(S)$ is $\textup{TH}_t$-exact.
1597: % \end{proof}
1598:
1599:
1600: \begin{remark} \label{rem:oddcycle} The theta-rank of $\I(S)$ could be
1601: much smaller than the upper bound in Remark~\ref{rem:k parallel
1602: translates}. Consider a $(2t+1)$-cycle $G$ and the set $S_G$ of
1603: characteristic vectors of its stable
1604: sets. Proposition~\ref{prop:2sos} shows that $\I(S_G)$ is
1605: $\textup{TH}_2$-exact. However, we need $t+1$ translates of the
1606: facet cut out by $\sum_{i=1}^{2t+1} x_i = t$ to cover $S_G$.
1607: \end{remark}
1608:
1609: In the rest of this section we derive various consequences of
1610: Theorem~\ref{thm:perfect}. Finite point sets with property (5) in
1611: Theorem~\ref{thm:perfect} have been studied in various contexts. In
1612: particular, Corollaries \ref{cor:perfect}, \ref{cor:perfectgraph} and
1613: \ref{cor:down-closed perfect} below were also observed independently
1614: by Greg Kuperberg, Raman Sanyal, Axel Werner and G{\"u}nter Ziegler
1615: (personal communication). In their work, $\textup{conv}(S)$ is called
1616: a {\bf 2-level polytope} when property (5) in
1617: Theorem~\ref{thm:perfect} holds.
1618:
1619: If $S$ is a finite subset of $\ZZ^n$ and $\mathcal L$ is the smallest
1620: lattice in $\ZZ^n$ containing $S$, then the lattice polytope
1621: $\textup{conv}(S)$ is said to be {\bf compressed} if every {\em
1622: reverse lexicographic} triangulation of the lattice points in
1623: $\textup{conv}(S)$ is {\em unimodular} with respect to $\mathcal L$.
1624: Compressed polytopes were introduced by Stanley
1625: \cite{StanleyCompressed}. Corollary~\ref{cor:perfect} (4) and Theorem
1626: 2.4 in \cite{Sullivant} (see also the references after Theorem 2.4 in
1627: \cite{Sullivant} for earlier citations of part or unpublished versions
1628: of this result), imply that a finite set $S \subset \RR^n$ is exact if
1629: and only if $\textup{conv}(S)$ is affinely equivalent to a compressed
1630: polytope.
1631:
1632: \begin{corollary} \label{cor:perfect} Let $S,S' \subset \RR^n$ be
1633: exact sets. Then
1634: \begin{enumerate}
1635: \item all points of $S$ are vertices of $\textup{conv}(S)$,
1636: \item the set of vertices of any face of $\textup{conv}(S)$ is again
1637: exact,
1638: \item the product $S \times S'$ is exact, and
1639: \item $\textup{conv}(S)$ is affinely equivalent to a $0/1$ polytope.
1640: \end{enumerate}
1641: \end{corollary}
1642:
1643: \begin{proof}
1644: The first three properties follow from Theorem~\ref{thm:perfect}
1645: (5). If the dimension of $\textup{conv}(S)$ is $d \,\,(\leq n)$,
1646: then $\textup{conv}(S)$ has at least $d$ non-parallel facets. If
1647: $\aa \cdot \xx \geq b$ cuts out a facet in this collection, then
1648: $\textup{conv}(S)$ is supported by both $\{ \xx \in \RR^n \,:\, \aa
1649: \cdot \xx = b\}$ and a parallel translate of it. Taking these two
1650: parallel hyperplanes from each of the $d$ facets gives a
1651: parallelepiped. By Theorem~\ref{thm:perfect}, $S$ is contained in
1652: the vertices of this parallelepiped intersected with the affine hull
1653: of $S$. This proves (4).
1654: \end{proof}
1655:
1656:
1657: By Corollary~\ref{cor:perfect} (4), it essentially suffices to look at
1658: subsets of $\{0,1\}^n$ to obtain all exact finite varieties in
1659: $\RR^n$. In $\RR^2$, the set of vertices of any $0/1$-polytope verify
1660: this property. In $\RR^3$ there are eight full-dimensional
1661: $0/1$-polytopes up to affine equivalence. In Figures~\ref{fig:perfect
1662: in r3} and \ref{fig:non-perfect in r3} the convex hulls of the exact
1663: and non-exact $0/1$ configurations in $\RR^3$ are shown.
1664:
1665: \begin{figure}
1666: \includegraphics[scale=0.35]{goodpoly.eps}
1667: \caption{Convex hulls of exact $0/1$ point sets in $\RR^3$.}
1668: \label{fig:perfect in r3}
1669: \end{figure}
1670:
1671: \begin{figure}
1672: \includegraphics[scale=0.35]{badpoly.eps}
1673: \caption{Convex hulls of non-exact $0/1$ point sets in $\RR^3$.}
1674: \label{fig:non-perfect in r3}
1675: \end{figure}
1676:
1677: \begin{example} \label{ex:perfect}
1678: The vertices of the following $0/1$-polytopes in $\RR^n$ are exact
1679: for every $n$: (1) hypercubes, (2) (regular) cross polytopes, (3)
1680: hypersimplices (includes simplices), (4) joins of $2$-level
1681: polytopes, and (5) stable set polytopes of perfect graphs on $n$
1682: vertices.
1683: \end{example}
1684:
1685: % \begin{example} \label{ex:notperfect}
1686: % In contrast to (5) in Example~\ref{ex:perfect}, the set $S$ of
1687: % vertices of $[0,1]^5$ that satisfy $\sum x_i = 1$ or $\sum x_i = 3$
1688: % is not exact. The convex hull of $S$ has 22 facets one of which is
1689: % defined by $- x_1 - x_2 - x_3 + x_4 + x_5 \leq 1$. The linear form
1690: % $-x_1 - x_2 - x_3 + x_4 + x_5$ takes values $-1,1
1691: % \,\textup{and}\,-3$ on the points of $S$ which proves that $S$ is
1692: % not exact.
1693: % \end{example}
1694:
1695: % \begin{example}
1696: % Exact zero-dimensional varieties may have convex hulls that are
1697: % neither simple nor simplicial. To see this, let $S_n$ be the
1698: % vertices of $P_n := \{ \xx \in \RR^n \,:\, 0 \leq x_i \leq 1
1699: % \,\,\forall \,\,i=1,\ldots, n, \,\,\, 1 \leq \sum x_i \leq 2 \}$.
1700: % Then $S$ has $n+ {n \choose 2}$ points and the $2n+2$ inequalities
1701: % in the above description of $P_n$ are precisely the facet
1702: % inequalities of $P_n$. Each point in $S$ is contained in $n+1$
1703: % facets. Further, $P_n$ has $n+1$ simplicial facets and $n+1$
1704: % non-simplicial facets, the latter with ${n \choose 2}$ vertices
1705: % each.
1706: % \end{example}
1707:
1708: \begin{theorem} \label{thm:2^n facets}
1709: If $S$ is a finite exact point set then $\textup{conv}(S)$ has at
1710: most $2^d$ facets and vertices, where $d=\dim \textup{conv}(S)$.
1711: Both bounds are sharp.
1712: \end{theorem}
1713:
1714: \begin{proof}
1715: The bound on the number of vertices is immediate by
1716: Corollary~\ref{cor:perfect} (4) and is achieved by $[0,1]^d$.
1717:
1718: For a polytope $P$ with an exact vertex set $S$, define a {\bf face
1719: pair} to be an unordered pair $(F_1,F_2)$ of proper faces of $P$
1720: such that $S \subseteq F_1 \cup F_2$ and $F_1$ and $F_2$ lie in
1721: parallel hyperplanes, or equivalently, there exists a linear form
1722: $h_{F_1,F_2}(\xx)$ such that $h_{F_1,F_2}(F_1)=0$ and
1723: $h_{F_1,F_2}(F_2)=1$. We will show that if $\textup{dim} \,\,P = d$
1724: then $P$ has at most $2^d-1$ face pairs and $2^d$ facets.
1725:
1726: If $d=1$, then an exact $S$ consists of two distinct points and $P$
1727: has two facets and one face pair as desired. Assume the result holds
1728: for $(d-1)$-polytopes with exact vertex sets and consider a
1729: $d$-polytope $P$ with exact vertex set $S$. Let $F$ be a facet of
1730: $P$ which by Theorem \ref{thm:perfect}, is in a face pair $(F,F')$
1731: of $P$. Since exactness does not depend on the affine embedding, we
1732: may assume that $P$ is full-dimensional and that $F$ spans the
1733: hyperplane $\{\xx \,:\, x_{d}=0 \}$, while $F'$ lies in $\{\xx \,:\,
1734: x_{d}=1\}$. By Corollary \ref{cor:perfect}, $F$ satisfies the
1735: induction hypothesis and so has at most $(2^{d-1}-1)$ face pairs.
1736: Any face pair of $P$ besides $(F,F')$ induces a face pair of $F$ by
1737: intersection with $F$, and every facet of $P$ is in a face pair of
1738: $P$ since $S$ is exact. The plan is to count how many face pairs of
1739: $P$ induce the same face pair of $F$ and the number of facets they
1740: contain.
1741:
1742: Fix a face pair $(F_1,F_2)$ of $F$, with associated linear form
1743: $h_{F_1,F_2}$ depending only on $x_1,\ldots,x_{d-1}$. Suppose
1744: $(F_1,F_2)$ is induced by a face pair of $P$ with associated linear
1745: form $H(\xx)$. Since $H$ and $h_{F_1,F_2}$ agree on every vertex of
1746: $F$, a facet of $P$, $H(\xx)=h_{F_1,F_2}(x_1,\ldots,x_{d-1})+cx_{d}$
1747: for some constant $c$.
1748:
1749: If $h_{F_1,F_2}(x_1,\ldots,x_{d-1})$ takes the same value $v$ on all
1750: of $F'$, then $H(F')=v+c = 0 \,\textup{or} \,1$ which implies that
1751: $c=-v$ or $c=1-v$. The two possibilities lead to the face pairs
1752: $(\textup{conv}(F_1 \cup F'), F_2)$ and $(\textup{conv}(F_2 \cup
1753: F'), F_1)$ of $P$. Each such pair contains at most one facet of
1754: $P$.
1755:
1756: If $h_{F_1,F_2}(x_1,\ldots,x_{d-1})$ takes more than one value on the
1757: vertices of $F'$, then these values must be $v$ and $v+1$ for some
1758: $v$ since $H$ takes values $0$ and $1$ on the vertices of $F'$. In
1759: that case, $c=-v$, so $H$ is unique and we get at most one face pair
1760: of $P$ inducing $(F_1,F_2)$. This pair will contain at most two
1761: facets of $P$.
1762:
1763: Since there are at most $2^{d-1}-1$ face pairs in $F$, they give us
1764: at most $2(2^{d-1}-1)$ face pairs and facets of $P$. Since we have
1765: not counted $(F,F')$ as a face pair of $P$, and $F$ and $F'$ as
1766: possible facets of $P$, we get the desired result. The bound on the
1767: number of facets is attained by cross-polytopes.
1768: \end{proof}
1769:
1770: \begin{remark}
1771: G{\"u}nter Ziegler has pointed out that our proof of
1772: Theorem~\ref{thm:2^n facets} can be refined to yield that $P$ (as
1773: used above) has $2^d-1$ face pairs if and only if it is a simplex
1774: and $2^d$ facets if and only if it is a regular cross-polytope.
1775: \end{remark}
1776:
1777:
1778:
1779: % \begin{remark}
1780: % G{\"u}nter Ziegler has pointed out the following refinements of
1781: % Theorem~\ref{thm:2^n facets} that follow from our proof. We state
1782: % these facts without proof using the notation used above.
1783: % \begin{enumerate}
1784: % \item As noted, the number of face pairs of $P$ is at most twice the
1785: % number of face pairs of $F$ plus one more coming from $(F,F')$.
1786: % Equality holds if and only if $P$ is a pyramid over $F$. In
1787: % particular, $P$ has at most $2^d-1$ face pairs with equality if and
1788: % only if $P$ is a simplex.
1789: % \item The proof also showed that the number of facets in $P$ is at
1790: % most twice the number of face pairs in $F$ plus two more from $F$
1791: % and $F'$ possibly. Equality occurs in situations such as when $P$
1792: % is a cube or a cross-polytope. Therefore, $P$ has at most $2^d$
1793: % facets and equality holds if and only if $P$ is affinely equivalent
1794: % to a regular cross-polytope.
1795: % \end{enumerate}
1796: % \end{remark}
1797: % Therefore, the number of face pairs of $P$ is at most twice the
1798: % number of face pairs of $F$ plus one more coming from $(F,F')$.
1799: % Equality holds if and only if $P$ is a pyramid over $F$. Indeed, if
1800: % $P$ is a pyramid over $F$, then $F'$ is a point and every face pair
1801: % $(F_1,F_2)$ in $F$ is induced by the two face pairs
1802: % $(\textup{conv}(F_1 \cup F'), F_2)$ and $(\textup{conv}(F_2 \cup
1803: % F'), F_1)$ of $P$. Conversely, if every face pair in $F$ is induced
1804: % by two face pairs in $P$, then since $F$ has $d-1$ non-parallel
1805: % facets, each of which is in a distinct face pair of $F$, there are
1806: % $d-1$ linear forms $h_{F_1,F_2}(x_1,\ldots,x_{d-1})$ with $d-1$
1807: % linearly independent normal vectors that take exactly one value each
1808: % on $F'$. This implies that $F'$ is a point and $P$ is a pyramid over
1809: % $F$. In particular, since there are at most $2^{d-1}-1$ face pairs
1810: % in $F$, $P$ has at most $2(2^{d-1}-1)+1 = 2^d-1$ face pairs with
1811: % equality if and only if $P$ is a simplex.
1812:
1813: % Also, the number of facets in $P$ is at most twice the number of
1814: % face pairs in $F$ plus two more from $F$ and $F'$ possibly.
1815: % Equality occurs in situations such as when $P$ is a cube or a
1816: % cross-polytope. Therefore, $P$ has at most $2(2^{d-1}-1)+2 = 2^d$
1817: % facets as claimed. Equality holds if and only if $P$ is affinely
1818: % equivalent to a regular cross-polytope. To see the ``only if''
1819: % direction of this assertion, recall that if $P$ has $2^d$ facets
1820: % then $F$ must have $2^{d-1}-1$ face pairs which implies that $F$ is
1821: % a $(d-1)$-simplex. Also, $F'$ must be facet of $P$ and it will also
1822: % have $2^{d-1}-1$ face pairs induced by each of the face pairs of $P$
1823: % inducing the $2^{d-1}-1$ face pairs of $F$. Therefore, $F'$ is also
1824: % a $(d-1)$-simplex. FINISH .... \marginpar{FINISH}
1825:
1826: Recall that Problem~\ref{prob:lovasz} was inspired by perfect graphs.
1827: Theorem~\ref{thm:perfect} adds to the characterizations of a perfect
1828: graph (c.f. Theorem~\ref{thm:knownforperfectgraphs}) as follows.
1829:
1830: \begin{corollary} \label{cor:perfectgraph}
1831: For a graph $G$, let $S_G$ denote the set of characteristic vectors
1832: of stable sets in $G$. Then the following are equivalent.
1833: \begin{enumerate}
1834: \item The graph $G$ is perfect.
1835: \item The stable set polytope, $\textup{STAB}(G)$, is a $2$-level
1836: polytope.
1837: % \item The set $S_G$ is exact.
1838: % \item The vanishing ideal $\I(S_G) = I_G$ is $(1,1)$-sos.
1839: % \item For each facet $F$ of $\textup{STAB}(G)$, $S_G$ is contained in
1840: % the union of $F$ and one other translate of the hyperplane spanned
1841: % by $F$. \marginpar{say 2-level polytope}
1842: \end{enumerate}
1843: \end{corollary}
1844:
1845: A polytope $P$ in $\RR^n_{\geq 0}$ is said to be {\bf down-closed} if
1846: for all $\vv \in P$ and $\vv' \in \RR^n_{\geq 0}$
1847: such that $v'_i \leq v_i$ for $i=1,\ldots,n$, $\vv' \in P$. For a
1848: graph $G$, $\textup{STAB}(G)$ is a down-closed $0/1$-polytope, and $G$
1849: is perfect if and only if the vertex set of $\textup{STAB}(G)$ is
1850: exact. We now prove that all down-closed $0/1$-polytopes with exact
1851: vertex sets are stable set polytopes of perfect graphs.
1852:
1853: \begin{theorem} \label{thm:down-closed perfect}
1854: Let $P \subseteq \RR^n$ be a down-closed $0/1$-polytope and $S$ be
1855: its set of vertices. Then $S$ is exact if and only if all facets of
1856: $P$ are either defined by non-negativity constraints on the
1857: variables or by an inequality of the form $\sum_{i \in I} x_i \leq
1858: 1$ for some $I \subseteq [n]$.
1859: \end{theorem}
1860:
1861: \begin{proof} If $P$ is not full-dimensional then since it is
1862: down-closed, it must be contained in a coordinate hyperplane $x_i =
1863: 0$ and the arguments below can be repeated in this lower-dimensional
1864: space. So we may assume that $P$ is $n$-dimensional. Then since $P$
1865: is down-closed, $S$ contains $\{{\bf 0},\ee_1, \ldots, \ee_n\}$.
1866:
1867: If all facets of $P$ are of the stated form, using that $S \subseteq
1868: \{0,1\}^n$, it is straight forward to check that $S$ is exact.
1869: % each $\ss \in S$ satisfies either $x_i = 0$ or $x_i =
1870: % 1$. Suppose $P$ has a facet inequality of the form $\sum_{i \in I}
1871: % x_i \leq 1$. Then since every $\ss \in S$ satisfies this inequality,
1872: % $S$ must be contained in $$\left\{ \xx \in \RR^n \,:\, \sum_{i \in
1873: % I} x_i = 1 \right\} \cup \left\{ \xx \in \RR^n \,:\, \sum_{i \in
1874: % I} x_i = 0 \right\}.$$
1875: % Therefore $S$ is exact.
1876:
1877: Now assume that $S$ is exact and $g(\xx) \geq 0$ is a facet
1878: inequality of $P$ that is not a non-negativity constraint. Then
1879: $g(\xx) := c - \sum_{i=1}^{n} a_i x_i \geq 0$ for some integers $c,
1880: a_1, \ldots, a_n$ with $c \neq 0$. Since ${\bf 0} \in S$ and $S$ is
1881: exact, we get that $g(\ss)$ equals $0$ or $c$ for all $\ss \in S$.
1882: Therefore, for all $i$, $g(\ee_i) = c-a_i$ equals $0$ or $c$, so
1883: $a_i$ is either $0$ or $c$. Dividing through by $c$, we get that the
1884: facet inequality $g(\xx) \geq 0$ is of the form $\sum_{i \in I} x_i
1885: \leq 1$ for some $I \subseteq [n]$.
1886: \end{proof}
1887:
1888: \begin{corollary} \label{cor:down-closed perfect}
1889: Let $P \subseteq \RR^n$ be a full-dimensional down-closed
1890: $0/1$-polytope and $S$ be its vertex set. Then $S$ is exact if and
1891: only if $P$ is the stable set polytope of a perfect graph.
1892: \end{corollary}
1893:
1894: \begin{proof}
1895: By Corollary~\ref{cor:perfectgraph} we only need to prove the
1896: ``only-if'' direction. Suppose $S$ is exact. Then by
1897: Theorem~\ref{thm:down-closed perfect}, all facet inequalities of $P$
1898: are either of the form $x_i \geq 0$ for some $i \in [n]$ or $\sum_{i
1899: \in I} x_i \leq 1$ for some $I \subseteq [n]$. Define the graph
1900: $G=([n],E)$ where $\{i,j\} \in E$ if and only if $\{i,j\} \subseteq
1901: I$ for some $I$ that indexes a facet inequality of $P$.
1902:
1903: We prove that $P = \textup{STAB}(G)$ and that $G$ is perfect. Let $K
1904: \subseteq [n]$ such that its characteristic vector $\chi^K \in S$.
1905: If there exists $i,j \in K$ such that $i,j \in I$ for some $I$ that
1906: indexes a facet inequality of $P$, then $1-\sum_{i \in I} x_i$ takes
1907: three different values when evaluated at the points ${\bf 0}, \ee_i,
1908: \chi^K$ in $S$ which contradicts that $S$ is exact.
1909: Therefore, $K$
1910: is a stable set of $G$ and $P \subseteq \textup{STAB}(G)$. If $K
1911: \subseteq [n]$ is a stable set of $G$ then, by construction, for
1912: every $I$ indexing a facet inequality of $P$, $\chi^K$ lies on
1913: either $\sum_{i \in I} x_i = 1$ or $\sum_{i \in I} x_i = 0$.
1914: Therefore $\chi^K \in P$ and $\textup{STAB}(G) \subseteq P$. Since
1915: all facet inequalities of $\textup{STAB}(G)$ are either
1916: non-negativities or clique inequalities, $G$ is perfect by
1917: \cite[Theorem~9.2.4 iii.]{GLS}.
1918: \end{proof}
1919:
1920:
1921: % Theorem~\ref{thm:perfect} suggests that for a finite set $S \subset
1922: % \RR^n$ we may want to consider the ``idempotent (polytope)
1923: % relaxation'' of $\conv(S)$ defined as
1924: % $$\textup{Ip}(S)=\{ \xx : g(\xx) \geq 0 \textrm{ for all } g \textrm{
1925: % linear and idempotent modulo }\I(S)\}.$$
1926: % Since all linear idempotent
1927: % polynomials are $1$-sos we immediately have that $\textup{TH}_1(\I(S))
1928: % \subseteq \textup{Ip}(S)$ and Theorem~\ref{thm:perfect} translates as
1929: % $\textup{TH}_1(\I(S))=\conv(S)$ if and only if
1930: % $\textup{Ip}(S)=\conv(S)$. For the stable set polytope, it is easy to
1931: % see that $\textup{Ip}(S_G)$ is simply the well-known polytope
1932: % relaxation $\textup{QSTAB}(G)$ of $\textup{STAB}(G)$, defined as
1933: % $$\textup{QSTAB}(G)=\left\{\xx \in \RR^n:x_i \geq 0; \sum_{i \in C}
1934: % x_i \leq 1, \textrm{ for all cliques } C \textrm{ of } G \right\}.$$
1935: % Thus in this case, Theorem~\ref{thm:perfect} recovers some of the
1936: % results in \cite{GLS} such as $\textup{TH}(G)=\textup{STAB}(G)$ if and
1937: % only if $\textup{QSTAB}=\textup{STAB}(G)$.
1938:
1939:
1940: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1941: \section{Arbitrary $\textup{TH}_1$-exact Ideals} \label{sec:structure
1942: arbitrary S}
1943:
1944: In this last section we describe $\textup{TH}_1(I)$ for an arbitrary
1945: (not necessarily real radical or zero-dimensional) ideal $I \subseteq
1946: \RR[\xx]$. The main structural result is
1947: Theorem~\ref{thm:intersection} which allows the construction of
1948: non-trivial high-dimensional $\textup{TH}_1$-exact ideals as in
1949: Example~\ref{ex:infinite perfect set}.
1950:
1951: In this study, the {\em convex quadrics} in $\RR[\xx]$ play a
1952: particularly important role. These are precisely the polynomials of
1953: degree two that can be written as $F(\xx)=\xx^t A \xx + \bb^t \xx +
1954: c$, where $A \not = 0$ is an $n \times n$ positive semidefinite matrix, $\bb
1955: \in \RR^n$ and $c \in \RR$. Note that every sum of squares of linear
1956: polynomials in $\RR[\xx]$ is a convex quadric.
1957:
1958:
1959: \begin{lemma}\label{lem:contained}
1960: For $I \subseteq \RR[\xx]$, $\textup{TH}_1(I) \neq \RR^n$ if and
1961: only if there exists some convex quadric $F \in I$.
1962: \end{lemma}
1963:
1964: \begin{proof}
1965: If $\textup{TH}_1(I) \neq \RR^n$, there exists a degree one polynomial $f$
1966: that is strictly positive on $\textup{TH}_1(I)$, hence $1$-sos modulo $I$.
1967: Then $f(\xx) \equiv g(\xx)$ mod $I$ for some $1$-sos $g(\xx) \not = 0$ and
1968: $g(\xx)-f(\xx) \in I$ is a convex quadric.
1969:
1970: Conversely, suppose $\xx^tA\xx + \bb^t \xx + c \in I$ with $A
1971: \succeq 0$. Then for any $\dd \in \RR^n$,
1972: $$(\xx+\dd)^tA(\xx+\dd) = \xx^tA\xx + 2\dd^tA\xx +\dd^tA\dd \equiv
1973: (2\dd^tA-\bb^t) \xx + \dd^tA\dd - c \mod I.$$
1974: Therefore, since
1975: $(\xx+\dd)^tA(\xx+\dd)$ is a sum of squares of linear polynomials,
1976: the linear polynomial $(2\dd^tA-\bb^t) \xx + \dd^tA\dd - c $ is
1977: $1$-sos mod $I$ and $\textup{TH}_1(I)$ must satisfy it. Since $\dd$
1978: can be chosen so that $(2\dd^tA-\bb^t) \neq 0$, $\textup{TH}_1(I)$
1979: is not trivial.
1980: \end{proof}
1981:
1982: % \begin{example}
1983: % For $S \subseteq \RR^2$, $\textup{TH}_1(\I(S)) \neq \RR^2$ if and
1984: % only if $S$ is contained in a pair of parallel lines, a parabola, or
1985: % an ellipse.
1986: % \end{example}
1987:
1988: % \begin{corollary}
1989: % If $\xx^tA\xx + \bb^t\xx + c \in I$ with $A
1990: % \succeq 0$ and of full rank then $\textup{TH}_1(I)$ is bounded.
1991: % \end{corollary}
1992:
1993: % \begin{proof}
1994: % In this case $\{2\dd^tA-\bb^t : \dd \in \RR^n\}=\RR^n$ and so there
1995: % are valid inequalities for $\textup{TH}_1(I)$ with all possible
1996: % elements of $\RR^n$ as normals.
1997: % \end{proof}
1998:
1999: \begin{lemma}\label{lem:theta_intersection}
2000: For an ideal $I \subseteq \RR[\xx]$, $\textup{TH}_1(I)=\bigcap
2001: \textup{TH}_1(\langle F \rangle)$, where $F$ varies over all convex
2002: quadrics in $I$.
2003: \end{lemma}
2004: \begin{proof}
2005: If $F \in I$ then $\langle F \rangle \subseteq I$. Also, if $f$
2006: is linear and $1$-sos mod $\langle F \rangle$ then it is also
2007: $1$-sos mod $I$. Therefore, $\textup{TH}_1(I) \subseteq
2008: \textup{TH}_1(\langle F \rangle)$.
2009:
2010: To prove the reverse inclusion, we need to show that if $f$ is a linear
2011: polynomial that is nonnegative on $\textup{TH}_1(I)$, it is also nonnegative
2012: on $\bigcap_{F \in I} \textup{TH}_1(\langle F \rangle)$, where $F$
2013: is a convex quadric. It suffices to show that whenever $f$ is linear
2014: and $1$-sos mod $I$, then there is a convex quadric $F \in I$ such
2015: that $f(\xx) \geq 0$ is valid for $\textup{TH}_1(\langle F
2016: \rangle)$, or equivalently that $f$ is $1$-sos mod $\langle F
2017: \rangle$. Since $f$ is $1$-sos mod $I$, there is a sum of squares of
2018: linear polynomials $g(\xx)$ such that $f(\xx) \equiv g(\xx)$ mod
2019: $I$. But $g$ is a convex quadric, hence so is $g(\xx)-f(\xx)$. Thus
2020: $f$ is $1$-sos mod the ideal $\langle g(\xx) - f(\xx) \rangle$ and
2021: we can take $F(\xx) = g(\xx)-f(\xx)$.
2022: \end{proof}
2023:
2024: \begin{lemma} \label{lemma:theta_convex}
2025: If $F(\xx)=\xx^tA\xx + \bb^t\xx + c$ with $A \succeq 0$, then
2026: $\textup{TH}_1(\langle F \rangle)=\textup{conv}(\V_{\RR}(F))$.
2027: \end{lemma}
2028: \begin{proof}
2029: We know that $\textup{conv}(\V_{\RR}(F)) \subseteq
2030: \textup{TH}_1(\langle F \rangle)$ and, since $F$ is convex,
2031: $\textup{conv}(\V_{\RR}(F))=\{\xx \in \RR^n \,:\, F(\xx)\leq 0\}$.
2032: Thus, if for every $\xx \in \V_{\RR}(F)$ $\textup{grad}F(\xx) \neq
2033: {\bf 0}$, then $\textup{conv}(\V_{\RR}(F))$ is supported by the
2034: tangent hyperplanes to $\V_{\RR}(F)$. In this case, to show that
2035: $\textup{TH}_1(\langle F \rangle) \subseteq
2036: \textup{conv}(\V_{\RR}(F))$, it suffices to prove that the defining
2037: (linear) polynomials of all tangent hyperplanes to $\V_{\RR}(F)$ are
2038: $1$-sos mod $\langle F \rangle$. The proof of the ``if'' direction
2039: of Lemma~\ref{lem:contained} shows that it would suffice to prove
2040: that a tangent hyperplane to $\V_{\RR}(F)$ has the form
2041: $(2\dd^tA-\bb^t) \xx + \dd^tA\dd - c = 0$, for some $\dd \in
2042: \RR^n$. The tangent at $\xx_0 \in \V_{\RR}(F)$ has equation
2043: $0=(2A\xx_0+\bb)^t(\xx-\xx_0)$ which can be rewritten as
2044: $$0= (2\xx_0^tA+\bb^t)\xx - 2\xx_0^tA\xx_0 -\bb^t\xx_0 =
2045: (2\xx_0^tA+\bb^t)\xx -\xx_0^tA\xx_0 + c,$$
2046: and so setting
2047: $\dd=-\xx_0$ gives the result.
2048:
2049: Suppose there is an $\xx_0$ such that $F(\xx_0) = 0$ and
2050: $\textup{grad} F(\xx_0) = {\bf 0}$. By translation we may assume
2051: that $\xx_0 = 0$, hence, $c = 0$ and $\bb = {\bf 0}$. Therefore $F =
2052: \xx^t A \xx = \sum h_i^2$ where the $h_i$ are linear. Since
2053: $\V_{\RR}(\langle F\rangle)=\V_{\RR}(\langle h_1,\ldots,h_m\rangle)$ it
2054: is enough to prove that all inequalities $\pm h_i \geq 0$ are valid
2055: for $\textup{TH}_1(\langle F \rangle)$. For any
2056: $\epsilon >0$ we have
2057: $$(\pm h_l + \epsilon)^2 + \sum_{i \not = l} h_i^2 = F \pm 2
2058: \epsilon h_l + \epsilon^2 \equiv 2 \epsilon(\pm h_l + \epsilon/2)
2059: \,\,\textup{mod} \,\, \langle F \rangle,$$
2060: so $\pm h_l + \epsilon/2$
2061: is $1$-sos mod $\langle F \rangle$ for all $l$ and all $\epsilon
2062: >0$. This implies that all the inequalities $\pm h_l + \epsilon/2
2063: \geq 0$ are valid for $\textup{TH}_1(\langle F \rangle)$, therefore
2064: so are the inequalities $\pm h_l \geq 0$.
2065: \end{proof}
2066:
2067: \begin{theorem} \label{thm:intersection} Let $I \subseteq \RR[\xx]$ be
2068: any ideal, then
2069: $$\textup{TH}_1(I)=\bigcap_{{F \in I}\atop{F \textrm{ convex
2070: quadric}}} \textup{conv}(\V_{\RR}(F)) = \bigcap_{{F \in
2071: I}\atop{F \textrm{ convex quadric}}} \{ \xx \in \RR^n : F(\xx)
2072: \leq 0\}.$$
2073: \end{theorem}
2074: \begin{proof}
2075: Immediate from Lemma \ref{lem:theta_intersection} and Lemma
2076: \ref{lemma:theta_convex}.
2077: \end{proof}
2078:
2079: \begin{example} \label{ex:infinite perfect set}
2080: Theorem~\ref{thm:intersection} shows that some non-principal ideals
2081: such as $I=\langle x^2-z,y^2-z \rangle \subseteq \RR[x,y,z]$ are
2082: $\textup{TH}_1$-exact. Since $\V_{\RR}(I)=\{(\pm t,\pm t,t^2): t
2083: \in \RR\}$, fixing the third coordinate we get the four points
2084: $(x,y,t^2)$ where $|x|=|y|=|t|$ which implies that
2085: $$\textup{conv}(\V_{\RR}(I))\supseteq \{(x,y,t^2): |x| \leq
2086: t, |y| \leq t, t \geq 0\}.$$ It is easy to see that the right hand
2087: side is equal to $\{(x,y,z): x^2 \leq z, y^2 \leq z\}$ which is
2088: exactly $\conv(\V_{\RR}(x^2-z)) \bigcap \conv(\V_{\RR}(y^2-z))$ and
2089: so contains $\textup{TH}_1(I)$ which contains
2090: $\textup{conv}(\V_{\RR}(I))$. So all inclusions must be equalities
2091: and $I$ is $\textup{TH}_1$-exact. This kind of reasoning allows us
2092: to construct non-trivial examples of $\textup{TH}_1$-exact ideals
2093: with high-dimensional varieties.
2094: \end{example}
2095:
2096: % \begin{example}
2097: % Theorem~\ref{thm:intersection} also gives us easy examples of
2098: % non-radical $\textup{TH}_1$-exact ideals. For example $I=\langle x^2
2099: % \rangle$ is $\textup{TH}_1$-exact since $x^2$ itself is a convex
2100: % quadric, or by the argument in Example~\ref{ex:conversefalse}.
2101: % \end{example}
2102:
2103: \begin{example} \label{ex:four points}
2104: Consider the set $S=\{(0,0),(1,0),(0,1),(2,2)\}$. Then the family of
2105: all quadratic curves in $\I(S)$ is
2106: $$a(x^2-x) + b(y^2-y) - (\frac{a+b}{2}) xy = (x,y) \left[
2107: \begin{array}{cc}
2108: a & -(\frac{a+b}{4}) \\
2109: -(\frac{a+b}{4}) & b
2110: \end{array}
2111: \right] \left( \begin{array}{c} x \\ y \end{array} \right) -ax -by.$$
2112: Since the case where both $a$ and $b$ are zero is trivial, we may
2113: normalize by setting $a+b=1$ and get the matrix in the quadratic to be
2114: $$\left[
2115: \begin{array}{cc}
2116: \lambda & -1/4 \\
2117: -1/4 & 1-\lambda
2118: \end{array}
2119: \right] $$
2120: with $\lambda \geq 0$. This matrix is positive semidefinite
2121: if and only if $\lambda(1-\lambda) - 1/16 \geq 0$, or equivalently, if
2122: and only if $\lambda \in [1/2 - \sqrt{3}/4, 1/2 + \sqrt{3}/4]$.
2123:
2124: This means that $(x,y) \in \textup{TH}_1(\I(S))$ if and only if, for
2125: all such $\lambda$,
2126: $$\lambda(x^2-x) + (1-\lambda)(y^2-y) -\frac{1}{2} xy \leq 0.$$
2127: Since the
2128: right-hand-side does not depend on $\lambda$, and the left-hand-side
2129: is a convex combination of $x^2-x$ and $y^2-y$, the inequality holds
2130: for every $\lambda \in [1/2 - \sqrt{3}/4, 1/2 + \sqrt{3}/4]$ if and
2131: only if it holds at the end points of the interval. Equivalently, if
2132: and only if
2133: $$\left(\frac{1}{2} - \frac{\sqrt{3}}{4}\right)(x^2-x) +
2134: \left(\frac{1}{2} + \frac{\sqrt{3}}{4}\right)(y^2-y) -\frac{1}{2}
2135: xy \leq 0,$$
2136: and
2137: $$\left(\frac{1}{2} + \frac{\sqrt{3}}{4}\right)(x^2-x) +
2138: \left(\frac{1}{2} - \frac{\sqrt{3}}{4}\right)(y^2-y) -\frac{1}{2}
2139: xy \leq 0.$$
2140: But this is just the intersection of the convex hull of the two
2141: curves obtained by turning the inequalities into equalities.
2142: Figure~\ref{fig:thetabody} shows this intersection.
2143:
2144: \begin{figure}
2145: \includegraphics[scale=0.35]{pic2.eps}
2146: \caption{Example~\ref{ex:four points}}
2147: \label{fig:thetabody}
2148: \end{figure}
2149: \end{example}
2150:
2151: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2152:
2153: \bibliographystyle{plain}
2154: \begin{thebibliography}{10}
2155:
2156: \bibitem{Barvinok}
2157: Alexander Barvinok.
2158: \newblock {\em A course in convexity}, volume~54 of {\em Graduate Studies in
2159: Mathematics}.
2160: \newblock American Mathematical Society, Providence, RI, 2002.
2161:
2162: \bibitem{BorweinLewis}
2163: Jonathan~M. Borwein and Adrian~S. Lewis.
2164: \newblock {\em Convex analysis and nonlinear optimization. Theory and
2165: examples}.
2166: \newblock CMS Books in Mathematics/Ouvrages de Math\'ematiques de la SMC, 3.
2167: Springer, New York, second edition, 2006.
2168:
2169: \bibitem{CLO}
2170: David Cox, John Little, and Donal O'Shea.
2171: \newblock {\em Ideals, {V}arieties and {A}lgorithms}.
2172: \newblock Springer-Verlag, New York, 1992.
2173:
2174: \bibitem{GoemansWilliamson}
2175: Michel~X. Goemans and David~P. Williamson.
2176: \newblock Improved approximation algorithms for maximum cut and satisfiability
2177: problems using semidefinite programming.
2178: \newblock {\em J. Assoc. Comput. Mach.}, 42(6):1115--1145, 1995.
2179:
2180: \bibitem{GLPT}
2181: Jo{\~a}o Gouveia, Monique Laurent, Pablo Parrilo, and Rekha Thomas.
2182: \newblock A new hierarchy of semidefinite programming relaxations for cycles in
2183: binary matroids and cuts in graphs.
2184: \newblock arXiv:0907.4518.
2185:
2186: \bibitem{GLS}
2187: Martin Gr{\"o}tschel, L{\'a}szl{\'o} Lov{\'a}sz, and Alexander Schrijver.
2188: \newblock {\em Geometric algorithms and combinatorial optimization}, volume~2
2189: of {\em Algorithms and Combinatorics}.
2190: \newblock Springer-Verlag, Berlin, second edition, 1993.
2191:
2192: \bibitem{HornJohnson}
2193: Roger~A. Horn and Charles~R. Johnson.
2194: \newblock {\em Matrix analysis}.
2195: \newblock Cambridge University Press, Cambridge, 1985.
2196:
2197: \bibitem{Lasserre1}
2198: Jean~B. Lasserre.
2199: \newblock Global optimization with polynomials and the problem of moments.
2200: \newblock {\em SIAM J. Optim.}, 11(3):796--817, 2001.
2201:
2202: \bibitem{Lasserre2}
2203: Jean~B. Lasserre.
2204: \newblock An explicit equivalent positive semidefinite program for nonlinear
2205: {$0$}-{$1$} programs.
2206: \newblock {\em SIAM J. Optim.}, 12(3):756--769, 2002.
2207:
2208: \bibitem{Lasserre3}
2209: Jean~B. Lasserre.
2210: \newblock Convex sets with semidefinite representation.
2211: \newblock {\em Math. Program.}, 120:457--477, 2009.
2212:
2213: \bibitem{LaLauRos}
2214: Jean~B. Lasserre, Monique Laurent, and Philipp Rostalski.
2215: \newblock Semidefinite characterization and computation of zero- dimensional
2216: real radical ideals.
2217: \newblock {\em Found. Comput. Math.}, 8(5):607--647, 2008.
2218:
2219: \bibitem{LaurentComparisonpaper}
2220: Monique Laurent.
2221: \newblock A comparison of the {S}herali-{A}dams, {L}ov\'asz-{S}chrijver, and
2222: {L}asserre relaxations for 0-1 programming.
2223: \newblock {\em Math. Oper. Res.}, 28(3):470--496, 2003.
2224:
2225: \bibitem{Laurent}
2226: Monique Laurent.
2227: \newblock Semidefinite representations for finite varieties.
2228: \newblock {\em Math. Program.}, 109(1, Ser. A):1--26, 2007.
2229:
2230: \bibitem{LaurentSosSurvey}
2231: Monique Laurent.
2232: \newblock Sums of squares, moment matrices and optimization over polynomials.
2233: \newblock In {\em Emerging Applications of Algebraic Geometry}, volume 149 of
2234: {\em IMA Volumes in Mathematics and its Applications}. Springer, 2009.
2235:
2236: \bibitem{LaurentRendl}
2237: Monique Laurent and Franz Rendl.
2238: \newblock Semidefinite programming and integer programming.
2239: \newblock In Karen Aardal, George Nemhauser, and Robert Weismantel, editors,
2240: {\em Handbook on Discrete Optimization}, pages 393--514. Elsevier B.V., 2005.
2241:
2242: \bibitem{ShannonCapacity}
2243: L{\'a}szl{\'o} Lov{\'a}sz.
2244: \newblock On the {S}hannon capacity of a graph.
2245: \newblock {\em IEEE Trans. Inform. Theory}, 25(1):1--7, 1979.
2246:
2247: \bibitem{LovaszStablesetsPolynomials}
2248: L{\'a}szl{\'o} Lov{\'a}sz.
2249: \newblock Stable sets and polynomials.
2250: \newblock {\em Discrete Math.}, 124(1-3):137--153, 1994.
2251: \newblock Graphs and combinatorics (Qawra, 1990).
2252:
2253: \bibitem{Lovasz}
2254: L{\'a}szl{\'o} Lov{\'a}sz.
2255: \newblock Semidefinite programs and combinatorial optimization.
2256: \newblock In {\em Recent advances in algorithms and combinatorics}, volume~11
2257: of {\em CMS Books Math./Ouvrages Math. SMC}, pages 137--194. Springer, New
2258: York, 2003.
2259:
2260: \bibitem{LovaszSchrijver91}
2261: L{\'a}szl{\'o} Lov{\'a}sz and Alexander Schrijver.
2262: \newblock Cones of matrices and set-functions and {$0$}-{$1$} optimization.
2263: \newblock {\em SIAM J. Optim.}, 1(2):166--190, 1991.
2264:
2265: \bibitem{MarshallBook}
2266: Murray Marshall.
2267: \newblock {\em Positive polynomials and sums of squares}, volume 146 of {\em
2268: Mathematical Surveys and Monographs}.
2269: \newblock American Math Society, Providence, RI, 2008.
2270:
2271: \bibitem{Parrilo:phd}
2272: Pablo~A. Parrilo.
2273: \newblock {\em Structured semidefinite programs and semialgebraic geometry
2274: methods in robustness and optimization}.
2275: \newblock PhD thesis, California Institute of Technology, May 2000.
2276: \newblock Available at
2277: \url{http://resolver.caltech.edu/CaltechETD:etd-05062004-055516}.
2278:
2279: \bibitem{parrilo}
2280: Pablo~A. Parrilo.
2281: \newblock An explicit construction of distinguished representations of
2282: polynomials nonnegative over finite sets.
2283: \newblock IfA Tech. Report AUT02-02, ETH Zurich, 2002.
2284:
2285: \bibitem{Parrilo:spr}
2286: Pablo~A. Parrilo.
2287: \newblock Semidefinite programming relaxations for semialgebraic problems.
2288: \newblock {\em Math. Prog.}, 96(2, Ser. B):293--320, 2003.
2289:
2290: \bibitem{SOSstructbook}
2291: Pablo~A. Parrilo.
2292: \newblock Exploiting algebraic structure in sum of squares programs.
2293: \newblock In D.~Henrion and A.~Garulli, editors, {\em Positive Polynomials in
2294: Control}, volume 312 of {\em Lecture Notes in Control and Information
2295: Sciences}. Springer, 2005.
2296:
2297: \bibitem{PowSchei}
2298: Victoria Powers and Claus Scheiderer.
2299: \newblock The moment problem for non-compact semialgebraic sets.
2300: \newblock {\em Adv. Geom.}, 1(1):71--88, 2001.
2301:
2302: \bibitem{Schoenebeck}
2303: Grant Schoenebeck.
2304: \newblock Linear level lasserre lower bounds for certain k-csps.
2305: \newblock In {\em FOCS}, pages 593--602. IEEE Computer Society, 2008.
2306:
2307: \bibitem{Sch}
2308: Alexander Schrijver.
2309: \newblock {\em Theory of Linear and Integer Programming}.
2310: \newblock Wiley-Interscience Series in Discrete Mathematics and Optimization,
2311: New York, 1986.
2312:
2313: \bibitem{SchrijverB}
2314: Alexander Schrijver.
2315: \newblock {\em Combinatorial optimization. {P}olyhedra and efficiency. {V}ol.
2316: {B}}, volume~24 of {\em Algorithms and Combinatorics}.
2317: \newblock Springer-Verlag, Berlin, 2003.
2318: \newblock Matroids, trees, stable sets, Chapters 39--69.
2319:
2320: \bibitem{BruceShepherd}
2321: F.~Bruce Shepherd.
2322: \newblock The theta body and imperfection.
2323: \newblock In {\em Perfect graphs}, Wiley-Intersci. Ser. Discrete Math. Optim.,
2324: pages 261--291. Wiley, Chichester, 2001.
2325:
2326: \bibitem{StanleyCompressed}
2327: Richard~P. Stanley.
2328: \newblock Decompositions of rational convex polytopes.
2329: \newblock {\em Ann. Discrete Math.}, 6:333--342, 1980.
2330: \newblock Combinatorial mathematics, optimal designs and their applications
2331: (Proc. Sympos. Combin. Math. and Optimal Design, Colorado State Univ., Fort
2332: Collins, Colo., 1978).
2333:
2334: \bibitem{Sullivant}
2335: Seth Sullivant.
2336: \newblock Compressed polytopes and statistical disclosure limitation.
2337: \newblock {\em Tohoku Math. J. (2)}, 58(3):433--445, 2006.
2338:
2339: \bibitem{BoydVandenberghe}
2340: Lieven Vandenberghe and Stephen Boyd.
2341: \newblock Semidefinite programming.
2342: \newblock {\em SIAM Rev.}, 38(1):49--95, 1996.
2343:
2344: \end{thebibliography}
2345:
2346: \end{document}
2347: