compile parker
[barvinok.git] / doc / implementation.tex
blob82eec2ce9714e3f2acd87e22c67f13e75f16da28
1 \section{Implementation details}
3 \subsection{An interior point of a polyhedron}
4 \label{s:interior}
6 We often need a point that lies in the interior of a polyhedron.
7 The function \ai[\tt]{inner\_point} implements the following algorithm.
8 Each polyhedron $P$ can be written as the sum of a polytope $P'$ and a cone $C$
9 (the \ai{recession cone} or \ai{characteristic cone} of $P$).
10 Adding a positive multiple of the sum of the extremal rays of $C$ to
11 the \ai{barycenter}
13 \frac 1 N \sum_i \vec v_i(\vec p)
15 of $P'$, where $N$ is the number of vertices, results in a point
16 in the interior of $P$.
18 \subsection{The integer points in the fundamental parallelepiped of a simple cone}
20 \label{s:fundamental}
22 This section is based on \shortciteN[Lemma 5.1]{Barvinok1992volume} and
23 \shortciteN{Koeppe2006experiments}.
25 \sindex{simple}{cone}
26 \sindex{open}{facet}
27 \sindex{open}{ray}
28 \sindex{explicit}{representation}
29 In this section we will deal exclusively with \ai{simple cone}s,
30 i.e. $d$-dimensional cones with $d$ extremal rays and $d$ facets.
31 \index{open facet}%
32 Some of the facets of these cones may be open.
33 Since we will mostly be dealing with cones in their
34 \ai{explicit representation}, we will have occasion to speak of
35 ``\ai{open ray}s'', by which we will mean that the facet not
36 containing the ray is open. (There is only one such facet because the cone
37 is simple.)
39 \sindex{fundamental}{parallelepiped}
40 \begin{definition}[Fundamental parallelepiped]
41 Let $K = \vec v + \poshull \lb\, \vec u_i \,\rb$ be
42 a closed (shifted) cone, then the \defindex{fundamental parallelepiped} $\Pi$
43 of $K$ is
45 \Pi = \vec v +
46 \lb\, \sum_i \alpha_i \vec u_i \mid 0 \leq \alpha_i < 1 \,\rb
49 If some of the rays $\vec u_i$ of $K$ are open, then the constraints on
50 the corresponding coefficient $\alpha_i$ are such that $0 < \alpha_i \le 1$.
51 \end{definition}
53 \begin{lemma}[Integer points in the fundamental parallelepiped of a simple cone]
54 \label{l:fundamental}
55 Let $K = \vec v + \poshull \lb\, \vec u_i \,\rb$ be a closed simple cone
56 and let $A$ be the matrix with the generators $\vec u_i$ of $K$
57 as rows.
58 Furthermore let $V A W^{-1} = S = \diag \vec s$ be the \indac{SNF} of $A$.
59 Then the integer points in the fundamental parallelepiped of $K$ are given
61 \begin{eqnarray}
62 \label{eq:parallelepiped}
63 \vec w^\T & = & \vec v^\T + \fractional{(\vec k^\T W - \vec v^\T) A^{-1}} A
65 \nonumber
66 & = &
67 \vec v^\T +
68 \sum_{i=1}^d
69 \fractional{\sps{\sum_{j=1}^d k_j \vec w^\T_j - \vec v^\T}{\vec u^*_i}}
70 \vec u_i^\T,
71 \end{eqnarray}
72 where $\vec u^*_i$ are the columns of $A^{-1}$ and $k_j \in \ZZ$ ranges
73 over $0 \le k_j < s_j$.
74 \end{lemma}
76 \begin{proof}
77 Since $0 \le \fractional{x} < 1$, it is clear that each such $\vec w$
78 lies inside the fundamental parallelepiped.
79 Furthermore,
80 \begin{eqnarray*}
81 \vec w^\T & = & \vec v^\T + \fractional{(\vec k^\T W - \vec v^\T) A^{-1}} A
83 & = &
84 \vec v^T +
85 \left(
86 (\vec k^\T W - \vec v^\T) A^{-1} - \floor{(\vec k^\T W - \vec v^\T) A^{-1}}
87 \right) A
89 & = &
90 \underbrace{\vec k^\T W\mathstrut}_{\in \ZZ^{1\times d}}
92 \underbrace{\floor{(\vec k^\T W - \vec v^\T) A^{-1}}}_{\in \ZZ^{1\times d}}
93 \underbrace{A\mathstrut}_{\in \ZZ^{d\times d}} \in \ZZ^{1\times d}.
94 \end{eqnarray*}
95 Finally, if two such $\vec w$ are equal, i.e., $\vec w_1 = \vec w_2$,
96 then
97 \begin{eqnarray*}
98 \vec 0^\T = \vec w_1^\T - \vec w_2^\T
99 & = & \vec k_1^\T W - \vec k_2^\T W + \vec p^\T A
101 & = & \left(\vec k_1^\T - \vec k_2^\T \right) W + \vec p^\T V^{-1} S W,
102 \end{eqnarray*}
103 with $\vec p \in \ZZ^d$,
104 or $\vec k_1 \equiv \vec k_2 \mod \vec s$, i.e., $\vec k_1 = \vec k_2$.
105 Since $\det S = \det A$, we obtain all points in the fundamental parallelepiped
106 by taking all $\vec k \in \ZZ^d$ satisfying $0 \le k_j < s_j$.
107 \end{proof}
109 If the cone $K$ is not closed then the coefficients of the open rays
110 should be in $(0,1]$ rather than in $[0,1)$.
111 In (\ref{eq:parallelepiped}),
112 we therefore need to replace the fractional part $\fractional{x} = x - \floor{x}$
113 by $\cractional{x} = x - \ceil{x-1}$ for the open rays.
115 \begin{figure}
116 \intercol=1.2cm
117 \begin{xy}
118 <\intercol,0pt>:<0pt,\intercol>::
119 \POS@i@={(0,-3),(0,0),(4,2),(4,-3)},{0*[grey]\xypolyline{*}}
120 \POS@i@={(0,-3),(0,0),(4,2)},{0*[|(2)]\xypolyline{}}
121 \POS(-1,0)\ar(4.5,0)
122 \POS(0,-3)\ar(0,3)
123 \POS(0,0)\ar@[|(3)](0,-1)
124 \POS(0,0)\ar@[|(3)](2,1)
125 \POS(0,-1)\ar@{--}@[|(2)](2,0)
126 \POS(2,1)\ar@{--}@[|(2)](2,0)
127 \POS(0,0)*{\bullet}
128 \POS(1,0)*{\bullet}
129 \end{xy}
130 \caption{The integer points in the fundamental parallelepiped of $K$}
131 \label{f:parallelepiped}
132 \end{figure}
134 \begin{example}
135 Let $K$ be the cone
137 K = \sm{0 \\ 0} + \poshull \lb\, \sm{2 \\ 1}, \sm{0 \\ -1} \,\rb
140 shown in Figure~\ref{f:parallelepiped}.
141 Then
143 A = \sm{2 & 1\\0 & -1} \qquad A^{-1} = \sm{1/2 & 1/2 \\ 0 & -1 }
147 \sm{1 & 0 \\ 1 & 1 } \sm{2 & 1\\0 & -1} = \sm{1 & 0 \\ 0 & 2} \sm{2 & 1 \\ 1 & 0}.
149 We have $\det A = \det S = 2$ and
150 $\vec k_1^\T = \sm{0 & 0}$ and $\vec k_2^\T = \sm{0 & 1}$.
151 Therefore,
153 \vec w_1^\T = \fractional{\sm{0 & 0} \sm{2 & 1 \\ 1 & 0} \sm{1/2 & 1/2 \\ 0 & -1 }}
154 \sm{2 & 1\\0 & -1} = \sm{0 & 0}
157 \begin{eqnarray*}
158 \vec w_2^\T & = &
159 \fractional{\sm{0 & 1} \sm{2 & 1 \\ 1 & 0} \sm{1/2 & 1/2 \\ 0 & -1 }}
160 \sm{2 & 1\\0 & -1}
162 & = &
163 \sm{1/2 & 1/2} \sm{2 & 1\\0 & -1} = \sm{1 & 0}.
164 \end{eqnarray*}
165 \end{example}
170 \subsection{Barvinok's decomposition of simple cones in primal space}
171 \label{s:decomposition}
173 As described by \shortciteN{DeLoera2003effective}, the first
174 implementation of Barvinok's counting algorithm applied
175 \ai{Barvinok's decomposition} \shortcite{Barvinok1994} in the \ai{dual space}.
176 \ai{Brion's polarization trick} \shortcite{Brion88} then ensures that you
177 do not need to worry about lower-dimensional faces in the decomposition.
178 Another way of avoiding the lower-dimensional faces, in the \ai{primal space},
179 is to perturb the vertex of the cone such that none of the lower-dimensional
180 face encountered contain any integer points \shortcite{Koeppe2006primal}.
181 In this section, we describe another technique that is based on allowing
182 some of the facets of the cone to be open.
184 The basic step in Barvinok's decomposition is to replace a
185 $d$-dimensional simple cone
186 $K = \poshull \lb\, \vec u_i \,\rb_{i=1}^d \subset \QQ^d$
187 by a signed sum of (at most) $d$ cones $K_j$
188 with a smaller determinant (in absolute value).
189 The cones are obtained by successively replacing each generator
190 of $K$ by an appropriately chosen
191 $\vec w = \sum_{i=1}^d \alpha_i \vec u_i$, i.e.,
192 \begin{equation}
193 \label{eq:K_j}
194 K_j =
195 \poshull \left(\lb\, \vec u_i \,\rb_{i=1}^d
196 \setminus \{\, \vec u_j \,\} \cup \{\, \vec w \,\}\right)
198 \end{equation}
199 To see that we can use these $K_j$ to perform a decomposition,
200 rearrange the $\vec u_i$ such that for all $1 \le i \le k$ we have
201 $\alpha_i < 0$ and for all $k+1 \le i \le d'$ we have $\alpha_i > 0$,
202 with $d - d'$ the number of zero $\alpha_i$.
203 We may assume $k < d'$; otherwise replace $\vec w \in B$ by
204 $-\vec w \in B$. We have
206 \vec w + \sum_{i=1}^k (-\alpha_i) \vec u_i =
207 \sum_{i=k+1}^{d'} \alpha_i \vec u_i
210 \begin{equation}
211 \label{eq:sub}
212 \sum_{i=0}^k \beta_i \vec u_i =
213 \sum_{i=k+1}^{d'} \alpha_i \vec u_i
215 \end{equation}
216 with $\vec u_0 = \vec w$, $\beta_0 = 1$ and $\beta_i = -\alpha_i > 0$
217 for $1 \le i \le k$. Any two $\vec u_j$ and $\vec u_l$ on the same side
218 of the equality are on opposite sides of the linear hull $H$ of
219 the other $\vec u_i$s since there exists a convex combination
220 of $\vec u_j$ and $\vec u_l$ on this hyperplane.
221 In particular, since $\alpha_j$ and $\alpha_l$ have the same sign,
222 we have
223 \begin{equation}
224 \label{eq:opposite}
225 \frac {\alpha_j}{\alpha_j+\alpha_l} \vec u_j
227 \frac {\alpha_l}{\alpha_j+\alpha_l} \vec u_l
228 \in H
229 \qquad\text{for $\alpha_i \alpha_l > 0$}
231 \end{equation}
232 The corresponding cones $K_j$ and $K_l$ (with $K_0 = K$)
233 therefore intersect in a common face $F \subset H$.
234 Let
236 K' :=
237 \poshull \left(\lb\, \vec u_i \,\rb_{i=1}^d \cup \{\, \vec w \,\}\right)
240 then any $\vec x \in K'$ lies both in some cone $K_i$ with
241 $0 \le i \le k$ and in some cone $K_i$ with $k+1 \le i \le d'$.
242 (Just subtract an appropriate multiple of Equation~(\ref{eq:sub}).)
243 The cones
244 $\{\, K_i \,\}_{i=0}^k$
246 $\{\, K_i \,\}_{i=k+1}^{d'}$
247 therefore both form a triangulation of $K'$ and hence
248 \begin{equation}
249 \label{eq:triangulations}
250 \indf{K'}
252 \indf{K} + \sum_{i=1}^k \indf{K_i} - \sum_{j\in J_1} \indf{F_j}
254 \sum_{i=k+1}^{d'} \indf{K_i} - \sum_{j\in J_2} \indf{F_j}
255 \end{equation}
257 \begin{equation}
258 \label{eq:decomposition}
259 \indf{K} = \sum_{i=1}^{d'} \varepsilon_i \indf{K_i} + \sum_j \delta_j \indf{F_j}
261 \end{equation}
262 with $\varepsilon_i = -1$ for $1 \le i \le k$,
263 $\varepsilon_i = 1$ for $k+1 \le i \le d'$,
264 $\delta_j \in \{ -1, 1 \}$ and $F_j$ some lower-dimensional faces.
265 Figure~\ref{fig:w} shows the possible configurations
266 in the case of a $3$-dimensional cone.
268 \begin{figure}
269 \intercol=0.48cm
270 \begin{center}
271 \begin{minipage}{0cm}
272 \begin{xy}
273 <\intercol,0pt>:<0pt,\intercol>::
275 \xybox{
276 \POS(-2,-1)="a"*+!U{+}
277 \POS(2,0)="b"*+!L{+}
278 \POS(0,2)="c"*+!D{+}
279 \POS(0,0)="w"*+!DR{\vec w}
280 \POS"a"\ar@{-}"b"
281 \POS"b"\ar@{-}"c"
282 \POS"c"\ar@{-}"a"
283 \POS"a"\ar@{--}"w"
284 \POS"b"\ar@{--}"w"
285 \POS"c"\ar@{--}"w"
286 }="a"
287 +R+(2,0)*!L
288 \xybox{
289 \POS(-2,-1)="a"*+!U{+}
290 \POS(2,0)="b"*+!L{-}
291 \POS(0,2)="c"*+!D{+}
292 \POS(-3,1)="w"*+!DR{\vec w}
293 \POS"a"\ar@{-}"b"
294 \POS"b"\ar@{-}"c"
295 \POS"c"\ar@{-}"a"
296 \POS"a"\ar@{--}"w"
297 \POS"b"\ar@{--}"w"
298 \POS"c"\ar@{--}"w"
299 }="b"
300 +R+(2,0)*!L
301 \xybox{
302 \POS(-2,-1)="a"*+!U{-}
303 \POS(2,0)="b"*+!U{+}
304 \POS(0,2)="c"*+!D{-}
305 \POS(5,-1)="w"*+!L{\vec w}
306 \POS"a"\ar@{-}"b"
307 \POS"b"\ar@{-}"c"
308 \POS"c"\ar@{-}"a"
309 \POS"a"\ar@{--}"w"
310 \POS"b"\ar@{--}"w"
311 \POS"c"\ar@{--}"w"
313 \POS"a"
314 +D-(0,1)*!U
315 \xybox{
316 \POS(-2,-1)="a"*+!U{0}
317 \POS(2,0)="b"*+!L{+}
318 \POS(0,2)="c"*+!D{+}
319 \POS(1,1)="w"*+!DL{\vec w}
320 \POS"a"\ar@{-}"b"
321 \POS"b"\ar@{-}"c"
322 \POS"c"\ar@{-}"a"
323 \POS"a"\ar@{--}"w"
325 \POS"b"
326 +DL-(0,1)*!UL
327 \xybox{
328 \POS(-2,-1)="a"*+!U{0}
329 \POS(2,0)="b"*+!U{+}
330 \POS(0,2)="c"*+!D{-}
331 \POS(4,-2)="w"*+!L{\vec w}
332 \POS"a"\ar@{-}"b"
333 \POS"b"\ar@{-}"c"
334 \POS"c"\ar@{-}"a"
335 \POS"a"\ar@{--}"w"
336 \POS"b"\ar@{--}"w"
338 \end{xy}
339 \end{minipage}
340 \end{center}
341 \caption[Possible locations of the vector $\vec w$ with respect to the rays
342 of a $3$-dimensional cone.]
343 {Possible locations of $\vec w$ with respect to the rays
344 of a $3$-dimensional cone. The figure shows a section of the cones.}
345 \label{fig:w}
346 \end{figure}
348 As explained above there are several ways of avoiding the lower-dimensional
349 faces in (\ref{eq:decomposition}). Here we will apply the following proposition.
350 \begin{proposition}[\shortciteN{Koeppe2008parametric}]
351 \label{p:inclusion-exclusion}
352 Let
353 \begin{equation}
354 \label{eq:full-source-identity}
355 \sum_{i\in {I_1}} \epsilon_i [P_i] + \sum_{i\in {I_2}} \delta_k [P_i] = 0
356 \end{equation}
357 be a (finite) linear identity of indicator functions of closed
358 polyhedra~$P_i\subseteq\QQ^d$, where the
359 polyhedra~$P_i$ with $i \in I_1$ are full-dimensional and those with $i \in I_2$
360 lower-dimensional. Let each closed polyhedron be given as
362 P_i = \left\{\, \vec x \mid \sp{b^*_{i,j}}{x} \ge \beta_{i,j} \text{
363 for $j\in J_i$}\,\right\}
366 Let $\vec y\in\QQ^d$ be a vector such that $\langle \vec b^*_{i,j}, \vec
367 y\rangle \neq 0$ for all $i\in I_1\cup I_2$, $j\in J_i$.
368 For each $i\in I_1$, we define the half-open polyhedron
369 \begin{equation}
370 \label{eq:half-open-by-y}
371 \begin{aligned}
372 \tilde P_i = \Bigl\{\, \vec x\in\QQ^d \mid {}&
373 \sp{b^*_{i,j}}{x} \ge \beta_{i,j}
374 \text{ for $j\in J_i$ with $\sp{b^*_{i,j}}{y} > 0$,} \\
375 & \sp{b^*_{i,j}}{x} > \beta_{i,j}
376 \text{ for $j\in J_i$ with $\sp{b^*_{i,j}}{y} < 0$} \,\Bigr\}.
377 \end{aligned}
378 \end{equation}
379 Then
380 \begin{equation}
381 \label{eq:target-identity}
382 \sum_{i\in I_1} \epsilon_i [\tilde P_i] = 0.
383 \end{equation}
384 \end{proposition}
385 When applying this proposition to (\ref{eq:decomposition}), we obtain
386 \begin{equation}
387 \label{eq:decomposition:2}
388 \indf{\tilde K} = \sum_{i=1}^{d'} \varepsilon_i \indf{\tilde K_i}
390 \end{equation}
391 where we start out
392 from a given $\tilde K$, which may be $K$ itself, i.e., a fully closed cone,
393 or the result of a previous application of the proposition, either through
394 a triangulation (Section~\ref{s:triangulation}) or a previous decomposition.
395 In either case, a suitable $\vec y$ is available, either as an interior
396 point of the cone or as the vector used in the previous application
397 (which may require a slight perturbation if it happens to lie on one of
398 the new facets of the cones $K_i$).
399 We are, however, free to construct a new $\vec y$ on each application
400 of the proposition.
401 In fact, we will not even construct such a vector explicitly, but
402 rather apply a set of rules that is equivalent to a valid choice of $\vec y$.
403 Below, we will present an ``intuitive'' motivation for these rules.
404 For a more algebraic, shorter, and arguably simpler motivation we
405 refer to \shortciteN{Koeppe2008parametric}.
407 The vector $\vec y$ has to satisfy $\sp{b^*_j}y > 0$ for normals $\vec b^*_j$
408 of closed facets and $\sp{b^*_j}y < 0$ for normals $\vec b^*_j$ of open facets of
409 $\tilde K$.
410 These constraints delineate a non-empty open cone $R$ from which
411 $\vec y$ should be selected. For some of the new facets of the cones
412 $\tilde K_j$, the cone $R$ will not be cut by the affine hull of the facet.
413 The closedness of these facets is therefore predetermined by $\tilde K$.
414 For the other facets, a choice will have to be made.
415 To be able to make the choice based on local information and without
416 computing an explicit vector $\vec y$, we use the following convention.
417 We first assign an arbitrary total order to the rays.
418 If (the affine hull of) a facet separates the two rays not on the facet $\vec u_i$
419 and $\vec u_j$, i.e., $\alpha_i \alpha_j > 0$ (\ref{eq:opposite}), then
420 we choose $\vec y$ to lie on the side of the smallest ray, according
421 to the chosen order.
422 That is, $\sp{{\tilde n}_{ij}}y > 0$, for
423 $\vec {\tilde n}_{ij}$ the normal of the facet pointing towards this smallest ray.
424 Otherwise, i.e., if $\alpha_i \alpha_j < 0$,
425 the interior of $K$ will lie on one side
426 of the facet and then we choose $\vec y$ to lie on the other side.
427 That is, $\sp{{\tilde n}_{ij}}y > 0$, for
428 $\vec {\tilde n}_{ij}$ the normal of the facet pointing away from the cone $K$.
429 Figure~\ref{fig:primal:examples} shows some example decompositions with
430 an explicitly marked $\vec y$.
432 \begin{figure}
433 \begin{align*}
434 \intercol=0.48cm
435 \begin{xy}
436 <\intercol,0pt>:<0pt,\intercol>::
437 \POS(-2,-1)="a"*+!U{+}
438 \POS(2,0)="b"*+!L{+}
439 \POS(0,2)="c"*+!D{+}
440 \POS"a"\ar@{-}@[|(3)]"b"
441 \POS"b"\ar@{-}@[|(3)]"c"
442 \POS"c"\ar@{-}@[|(3)]"a"
443 \POS(0.3,0.6)*{\bullet},*+!L{\vec y}
444 \end{xy}
446 \intercol=0.48cm
447 \begin{xy}
448 <\intercol,0pt>:<0pt,\intercol>::
449 \POS(2,0)="b"*+!L{+}
450 \POS(0,2)="c"*+!D{+}
451 \POS(0,0)="w"*+!DR{\vec w}
452 \POS"b"\ar@{-}@[|(3)]"c"
453 \POS"b"\ar@{-}@[|(3)]"w"
454 \POS"c"\ar@{-}@[|(3)]"w"
455 \POS(0.3,0.6)*{\bullet},*+!L{\vec y}
456 \end{xy}
458 \begin{xy}
459 <\intercol,0pt>:<0pt,\intercol>::
460 \POS(-2,-1)="a"*+!U{+}
461 \POS(0,2)="c"*+!D{+}
462 \POS(0,0)="w"*+!DR{\vec w}
463 \POS"c"\ar@{-}@[|(3)]"a"
464 \POS"a"\ar@{-}@[|(3)]"w"
465 \POS"c"\ar@{--}@[|(3)]"w"
466 \POS(0.3,0.6)*{\bullet},*+!L{\vec y}
467 \end{xy}
469 \begin{xy}
470 <\intercol,0pt>:<0pt,\intercol>::
471 \POS(-2,-1)="a"*+!U{+}
472 \POS(2,0)="b"*+!L{+}
473 \POS(0,0)="w"*+!DR{\vec w}
474 \POS"a"\ar@{-}@[|(3)]"b"
475 \POS"a"\ar@{--}@[|(3)]"w"
476 \POS"b"\ar@{--}@[|(3)]"w"
477 \POS(0.3,0.6)*{\bullet},*+!L{\vec y}
478 \end{xy}
480 \intercol=0.48cm
481 \begin{xy}
482 <\intercol,0pt>:<0pt,\intercol>::
483 \POS(-2,-1)="a"*+!U{+}
484 \POS(2,0)="b"*+!L{+}
485 \POS(0,2)="c"*+!D{+}
486 \POS"a"\ar@{--}@[|(3)]"b"
487 \POS"b"\ar@{-}@[|(3)]"c"
488 \POS"c"\ar@{--}@[|(3)]"a"
489 \POS(-2.5,-1.5)*{\bullet},*+!U{\vec y}
490 \end{xy}
492 \intercol=0.48cm
493 \begin{xy}
494 <\intercol,0pt>:<0pt,\intercol>::
495 \POS(2,0)="b"*+!L{+}
496 \POS(0,2)="c"*+!D{+}
497 \POS(0,0)="w"*+!DR{\vec w}
498 \POS"b"\ar@{-}@[|(3)]"c"
499 \POS"b"\ar@{--}@[|(3)]"w"
500 \POS"c"\ar@{--}@[|(3)]"w"
501 \POS(-2.5,-1.5)*{\bullet},*+!U{\vec y}
502 \end{xy}
504 \begin{xy}
505 <\intercol,0pt>:<0pt,\intercol>::
506 \POS(-2,-1)="a"*+!U{+}
507 \POS(0,2)="c"*+!D{+}
508 \POS(0,0)="w"*+!DR{\vec w}
509 \POS"c"\ar@{--}@[|(3)]"a"
510 \POS"a"\ar@{--}@[|(3)]"w"
511 \POS"c"\ar@{-}@[|(3)]"w"
512 \POS(-2.5,-1.5)*{\bullet},*+!U{\vec y}
513 \end{xy}
515 \begin{xy}
516 <\intercol,0pt>:<0pt,\intercol>::
517 \POS(-2,-1)="a"*+!U{+}
518 \POS(2,0)="b"*+!L{+}
519 \POS(0,0)="w"*+!DR{\vec w}
520 \POS"a"\ar@{--}@[|(3)]"b"
521 \POS"a"\ar@{-}@[|(3)]"w"
522 \POS"b"\ar@{-}@[|(3)]"w"
523 \POS(-2.5,-1.5)*{\bullet},*+!U{\vec y}
524 \end{xy}
526 \intercol=0.48cm
527 \begin{xy}
528 <\intercol,0pt>:<0pt,\intercol>::
529 \POS(-2,-1)="a"*+!U{+}
530 \POS(2,0)="b"*+!L{+}
531 \POS(0,2)="c"*+!D{+}
532 \POS"a"\ar@{--}@[|(3)]"b"
533 \POS"b"\ar@{-}@[|(3)]"c"
534 \POS"c"\ar@{-}@[|(3)]"a"
535 \POS(1,-1.5)*{\bullet},*+!L{\vec y}
536 \end{xy}
538 \intercol=0.48cm
539 \begin{xy}
540 <\intercol,0pt>:<0pt,\intercol>::
541 \POS(2,0)="b"*+!L{+}
542 \POS(0,2)="c"*+!D{+}
543 \POS(0,0)="w"*+!DR{\vec w}
544 \POS"b"\ar@{-}@[|(3)]"c"
545 \POS"b"\ar@{--}@[|(3)]"w"
546 \POS"c"\ar@{-}@[|(3)]"w"
547 \POS(1,-1.5)*{\bullet},*+!L{\vec y}
548 \end{xy}
550 \begin{xy}
551 <\intercol,0pt>:<0pt,\intercol>::
552 \POS(-2,-1)="a"*+!U{+}
553 \POS(0,2)="c"*+!D{+}
554 \POS(0,0)="w"*+!DR{\vec w}
555 \POS"c"\ar@{-}@[|(3)]"a"
556 \POS"a"\ar@{--}@[|(3)]"w"
557 \POS"c"\ar@{--}@[|(3)]"w"
558 \POS(1,-1.5)*{\bullet},*+!L{\vec y}
559 \end{xy}
561 \begin{xy}
562 <\intercol,0pt>:<0pt,\intercol>::
563 \POS(-2,-1)="a"*+!U{+}
564 \POS(2,0)="b"*+!L{+}
565 \POS(0,0)="w"*+!DR{\vec w}
566 \POS"a"\ar@{--}@[|(3)]"b"
567 \POS"a"\ar@{-}@[|(3)]"w"
568 \POS"b"\ar@{-}@[|(3)]"w"
569 \POS(1,-1.5)*{\bullet},*+!L{\vec y}
570 \end{xy}
572 \intercol=0.48cm
573 \begin{xy}
574 <\intercol,0pt>:<0pt,\intercol>::
575 \POS(-2,-1)="a"*+!U{0}
576 \POS(2,0)="b"*+!L{+}
577 \POS(0,2)="c"*+!D{+}
578 \POS"a"\ar@{-}@[|(3)]"b"
579 \POS"b"\ar@{-}@[|(3)]"c"
580 \POS"c"\ar@{-}@[|(3)]"a"
581 \POS(1,0.2)*{\bullet},*+!R{\vec y}
582 \end{xy}
584 \intercol=0.48cm
585 \begin{xy}
586 <\intercol,0pt>:<0pt,\intercol>::
587 \POS(-2,-1)="a"*+!U{0}
588 \POS(2,0)="b"*+!L{+}
589 \POS(1,1)="w"*+!DL{\vec w}
590 \POS"a"\ar@{-}@[|(3)]"b"
591 \POS"a"\ar@{-}@[|(3)]"w"
592 \POS"b"\ar@{-}@[|(3)]"w"
593 \POS(1,0.2)*{\bullet},*+!R{\vec y}
594 \end{xy}
596 \begin{xy}
597 <\intercol,0pt>:<0pt,\intercol>::
598 \POS(-2,-1)="a"*+!U{0}
599 \POS(0,2)="c"*+!D{+}
600 \POS(1,1)="w"*+!DL{\vec w}
601 \POS"c"\ar@{-}@[|(3)]"a"
602 \POS"a"\ar@{--}@[|(3)]"w"
603 \POS"c"\ar@{-}@[|(3)]"w"
604 \POS(1,0.2)*{\bullet},*+!R{\vec y}
605 \end{xy}
607 \intercol=0.48cm
608 \begin{xy}
609 <\intercol,0pt>:<0pt,\intercol>::
610 \POS(-2,-1)="a"*+!U{0}
611 \POS(2,0)="b"*+!U{+}
612 \POS(0,2)="c"*+!D{-}
613 \POS"a"\ar@{-}@[|(3)]"b"
614 \POS"b"\ar@{--}@[|(3)]"c"
615 \POS"c"\ar@{-}@[|(3)]"a"
616 \POS(1.5,1.5)*{\bullet},*+!D{\vec y}
617 \end{xy}
619 \intercol=0.48cm
620 \begin{xy}
621 <\intercol,0pt>:<0pt,\intercol>::
622 \POS(-2,-1)="a"*+!U{0}
623 \POS(0,2)="c"*+!D{-}
624 \POS(4,-2)="w"*+!L{\vec w}
625 \POS"c"\ar@{-}@[|(3)]"a"
626 \POS"a"\ar@{-}@[|(3)]"w"
627 \POS"c"\ar@{--}@[|(3)]"w"
628 \POS(1.5,1.5)*{\bullet},*+!D{\vec y}
629 \end{xy}
631 \begin{xy}
632 <\intercol,0pt>:<0pt,\intercol>::
633 \POS(-2,-1)="a"*+!U{0}
634 \POS(2,0)="b"*+!U{+}
635 \POS(4,-2)="w"*+!L{\vec w}
636 \POS"a"\ar@{--}@[|(3)]"b"
637 \POS"a"\ar@{-}@[|(3)]"w"
638 \POS"b"\ar@{--}@[|(3)]"w"
639 \POS(1.5,1.5)*{\bullet},*+!D{\vec y}
640 \end{xy}
642 \intercol=0.48cm
643 \begin{xy}
644 <\intercol,0pt>:<0pt,\intercol>::
645 \POS(-2,-1)="a"*+!U{0}
646 \POS(2,0)="b"*+!U{+}
647 \POS(0,2)="c"*+!D{-}
648 \POS"a"\ar@{--}@[|(3)]"b"
649 \POS"b"\ar@{--}@[|(3)]"c"
650 \POS"c"\ar@{-}@[|(3)]"a"
651 \POS(4.7,-2.5)*{\bullet},*+!R{\vec y}
652 \end{xy}
654 \intercol=0.48cm
655 \begin{xy}
656 <\intercol,0pt>:<0pt,\intercol>::
657 \POS(-2,-1)="a"*+!U{0}
658 \POS(0,2)="c"*+!D{-}
659 \POS(4,-2)="w"*+!L{\vec w}
660 \POS"c"\ar@{-}@[|(3)]"a"
661 \POS"a"\ar@{--}@[|(3)]"w"
662 \POS"c"\ar@{--}@[|(3)]"w"
663 \POS(4.7,-2.5)*{\bullet},*+!R{\vec y}
664 \end{xy}
666 \begin{xy}
667 <\intercol,0pt>:<0pt,\intercol>::
668 \POS(-2,-1)="a"*+!U{0}
669 \POS(2,0)="b"*+!U{+}
670 \POS(4,-2)="w"*+!L{\vec w}
671 \POS"a"\ar@{-}@[|(3)]"b"
672 \POS"a"\ar@{--}@[|(3)]"w"
673 \POS"b"\ar@{--}@[|(3)]"w"
674 \POS(4.7,-2.5)*{\bullet},*+!R{\vec y}
675 \end{xy}
676 \end{align*}
677 \caption{Examples of decompositions in primal space.}
678 \label{fig:primal:examples}
679 \end{figure}
681 To see that there is a $\vec y$ satisfying the above constraints,
682 we need to show that $R \cap S$ is non-empty, with
683 $S = \{ \vec y \mid \sp{{\tilde n}_{i_kj_k}}y > 0 \text{ for all $k$}\}$.
684 It will be easier to show this set is non-empty when the $\vec u_i$ form
685 an orthogonal basis. Applying a non-singular linear transformation $T$
686 does not change the decomposition of $\vec w$ in terms of the $\vec u_i$ (i.e., the
687 $\alpha_i$ remain unchanged), nor does this change
688 any of the scalar products in the constraints that define $R \cap S$
689 (the normals are transformed by $\left(T^{-1}\right)^\T$).
690 Finding a vector $\vec y \in T(R \cap S)$ ensures that
691 $T^{-1}(\vec y) \in R \cap S$.
692 Without loss of generality, we can therefore assume for the purpose of
693 showing that $R \cap S$ is non-empty that
694 the $\vec u_i$ indeed form an orthogonal basis.
696 In the orthogonal basis, we have $\vec b_i^* = \vec u_i$
697 and the corresponding inward normal $\vec N_i$ is either
698 $\vec u_i$ or $-\vec u_i$.
699 Furthermore, each normal of a facet of $S$ of the first type is of the
700 form $\vec {\tilde n}_{i_kj_k} = a_k \vec u_{i_k} - b_k \vec u_{j_k}$, with
701 $a_k, b_k > 0$ and ${i_k} < {j_k}$,
702 while for the second type each normal is of the form
703 $\vec {\tilde n}_{i_kj_k} = -a_k \vec u_{i_k} - b_k \vec u_{j_k}$, with
704 $a_k, b_k > 0$.
705 If $\vec {\tilde n}_{i_kj_k} = a_k \vec u_{i_k} - b_k \vec u_{j_k}$
706 is the normal of a facet of $S$
707 then either
708 $(\vec N_{i_k}, \vec N_{j_k}) = (\vec u_{i_k}, \vec u_{j_k})$
710 $(\vec N_{i_k}, \vec N_{j_k}) = (-\vec u_{i_k}, -\vec u_{j_k})$.
711 Otherwise, the facet would not cut $R$.
712 Similarly,
713 if $\vec {\tilde n}_{i_kj_k} = -a_k \vec u_{i_k} - b_k \vec u_{j_k}$
714 is the normal of a facet of $S$
715 then either
716 $(\vec N_{i_k}, \vec N_{j_k}) = (\vec u_{i_k}, -\vec u_{j_k})$
718 $(\vec N_{i_k}, \vec N_{j_k}) = (-\vec u_{i_k}, \vec u_{j_k})$.
719 Assume now that $R \cap S$ is empty, then there exist
720 $\lambda_k, \mu_i \ge 0$ not all zero
721 such that
722 $\sum_k \lambda_k \vec {\tilde n}_{i_kj_k} + \sum_l \mu_i \vec N_i = \vec 0$.
723 Assume $\lambda_k > 0$ for some facet of the first type.
724 If $\vec N_{j_k} = -\vec u_{j_k}$, then $-b_k$ can only be canceled
725 by another facet $k'$ of the first type with $j_k = i_{k'}$, but then
726 also $\vec N_{j_{k'}} = -\vec u_{j_{k'}}$. Since the $j_k$ are strictly
727 increasing, this sequence has to stop with a strictly positive coefficient
728 for the largest $\vec u_{j_k}$ in this sequence.
729 If, on the other hand, $\vec N_{i_k} = \vec u_{i_k}$, then $a_k$ can only
730 be canceled by the normal of a facet $k'$ of the second kind
731 with $i_k = j_{k'}$, but then
732 $\vec N_{i_{k'}} = -\vec u_{i_{k'}}$ and we return to the first case.
733 Finally, if $\lambda_k > 0$ only for normals of facets of the second type,
734 then either $\vec N_{i_k} = -\vec u_{i_k}$ or $\vec N_{j_k} = -\vec u_{j_k}$
735 and so the coefficient of one of these basis vectors will be strictly
736 negative.
737 That is, the sum of the normals will never be zero and
738 the set $R \cap S$ is non-empty.
740 For each ray $\vec u_j$ of cone $K_i$, i.e., the cone with $\vec u_i$ replaced
741 by $\vec w$, we now need to determine whether the facet not containing this
742 ray is closed or not. We denote the (inward) normal of this cone by
743 $\vec n_{ij}$. Note that cone $K_j$ (if it appears in (\ref{eq:triangulations}),
744 i.e., $\alpha_j \ne 0$) has the same facet opposite $\vec u_i$
745 and its normal $\vec n_{ji}$ will be equal to either $\vec n_{ij}$ or
746 $-\vec n_{ij}$, depending on whether we are dealing with an ``external'' facet,
747 i.e., a facet of $K'$, or an ``internal'' facet.
748 If, on the other hand, $\alpha_j = 0$, then $\vec n_{ij} = \vec n_{0j}$.
749 If $\sp{n_{ij}}y > 0$, then the facet is closed.
750 Otherwise it is open.
751 It follows that the two (or more) occurrences of external facets are either all open
752 or all closed, while for internal facets, exactly one is closed.
754 First consider the facet not containing $\vec u_0 = \vec w$.
755 If $\alpha_i > 0$, then $\vec u_i$ and $\vec w$ are on the same side of the facet
756 and so $\vec n_{i0} = \vec n_{0i}$. Otherwise, $\vec n_{i0} = -\vec n_{i0}$.
757 Second, if $\alpha_j = 0$, then replacing $\vec u_i$ by $\vec w$ does not
758 change the affine hull of the facet and so $\vec n_{ij} = \vec n_{0j}$.
759 Now consider the case that $\alpha_i \alpha_j < 0$, i.e., $\vec u_i$
760 and $\vec u_j$ are on the same side of the hyperplane through the other rays.
761 If we project $\vec u_i$, $\vec u_j$ and $\vec w$ onto a plane orthogonal
762 to the ridge through the other rays, then the possible locations of $\vec w$
763 with respect to $\vec u_i$ and $\vec u_j$ are shown in Figure~\ref{fig:w:same}.
764 If both $\vec n_{0i}$ and $\vec n_{0j}$ are closed then $\vec y$ lies in region~1
765 and therefore $\vec n_{ij}$ (as well as $\vec n_{ji}$) is closed too.
766 Similarly, if both $\vec n_{0i}$ and $\vec n_{0j}$ are open then so is
767 $\vec n_{ij}$. If only one of the facets is closed, then, as explained above,
768 we choose $\vec n_{ij}$ to be open, i.e., we take $\vec y$ to lie in region~3
769 or~5.
770 Figure~\ref{fig:w:opposite} shows the possible configurations
771 for the case that $\alpha_i \alpha_j > 0$.
772 If exactly one of $\vec n_{0i}$ and $\vec n_{0j}$ is closed, then
773 $\vec y$ lies in region~3 or region~5 and therefore $\vec n_{ij}$ is closed iff
774 $\vec n_{0j}$ is closed.
775 Otherwise, as explained above, we choose $\vec n_{ij}$ to be closed if $i < j$.
777 \begin{figure}
778 \intercol=0.7cm
779 \begin{minipage}{0cm}
780 \begin{xy}
781 <\intercol,0pt>:<0pt,\intercol>::
782 \POS(-4,0)="a"\ar@[|(2)]@{-}(4,0)
783 \POS?(0)/\intercol/="b"\POS(2,0)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0i}}(0,0.75)}
784 \POS(2,0)*{\bullet},*+!U{\vec u_j}
785 \POS(-2,-3)="a"\ar@[|(2)]@{-}(2,3)
786 \POS?(0)/\intercol/="b"\POS(1,1.5)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0j}}(0,-0.75)}
787 \POS(1,1.5)*{\bullet},*+!R{\vec u_i}
788 \POS(-2,3)="a"\ar@{-}(2,-3)
789 \POS?(0)/\intercol/="b"\POS(1.5,-2.25)
790 *\xybox{"b"-"a":(0,0)\ar_{\vec n_{ji}}^{\vec n_{ij}}(0,+0.75)}
791 \POS(1.5,-2.25)*{\bullet},*+!R{\vec u_0 = \vec w}
792 \POS(0,0)*{\bullet}
793 \POS(3,1.5)*+[o][F]{\scriptstyle 1}
794 \POS(0,2.5)*+[o][F]{\scriptstyle 2}
795 \POS(-3,1.5)*+[o][F]{\scriptstyle 3}
796 \POS(-3,-1.5)*+[o][F]{\scriptstyle 4}
797 \POS(0,-3)*+[o][F]{\scriptstyle 5}
798 \POS(3,-1.5)*+[o][F]{\scriptstyle 6}
799 \end{xy}
800 \end{minipage}
801 \begin{minipage}{0cm}
802 \begin{xy}
803 <\intercol,0pt>:<0pt,\intercol>::
804 \POS(-4,0)="a"\ar@[|(2)]@{-}(4,0)
805 \POS?(0)/\intercol/="b"\POS(2,0)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0i}}(0,0.75)}
806 \POS(2,0)*{\bullet},*+!U{\vec u_j}
807 \POS(-2,-3)="a"\ar@[|(2)]@{-}(2,3)
808 \POS?(0)/\intercol/="b"\POS(1,1.5)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0j}}(0,-0.75)}
809 \POS(1,1.5)*{\bullet},*+!R{\vec u_i}
810 \POS(-2,3)="a"\ar@{-}(2,-3)
811 \POS?(0)/\intercol/="b"\POS(-1.5,2.25)
812 *\xybox{"b"-"a":(0,0)\ar_{\vec n_{ji}}^{\vec n_{ij}}(0,+0.75)}
813 \POS(-1.5,2.25)*{\bullet},*+!R{\vec u_0 = \vec w}
814 \POS(0,0)*{\bullet}
815 \POS(3,1.5)*+[o][F]{\scriptstyle 1}
816 \POS(0,2.5)*+[o][F]{\scriptstyle 2}
817 \POS(-3,1.5)*+[o][F]{\scriptstyle 3}
818 \POS(-3,-1.5)*+[o][F]{\scriptstyle 4}
819 \POS(0,-3)*+[o][F]{\scriptstyle 5}
820 \POS(3,-1.5)*+[o][F]{\scriptstyle 6}
821 \end{xy}
822 \end{minipage}
823 \caption{Possible locations of $\vec w$ with respect to $\vec u_i$ and
824 $\vec u_j$, projected onto a plane orthogonal to the other rays, when
825 $\alpha_i \alpha_j < 0$.}
826 \label{fig:w:same}
827 \end{figure}
829 \begin{figure}
830 \intercol=0.7cm
831 \begin{minipage}{0cm}
832 \begin{xy}
833 <\intercol,0pt>:<0pt,\intercol>::
834 \POS(-4,0)="a"\ar@[|(2)]@{-}(4,0)
835 \POS?(0)/\intercol/="b"\POS(2,0)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0i}}(0,0.75)}
836 \POS(2,0)*{\bullet},*+!U{\vec u_j}
837 \POS(-2,3)="a"\ar@[|(2)]@{-}(2,-3)
838 \POS?(0)/\intercol/="b"\POS(-1,1.5)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0j}}(0,+0.75)}
839 \POS(-1,1.5)*{\bullet},*+!R{\vec u_i}
840 \POS(-2,-3)="a"\ar@{-}(2,3)
841 \POS?(0)/\intercol/="b"\POS(1.5,2.25)
842 *\xybox{"b"-"a":(0,0)\ar^{\vec n_{ji}}(0,+0.75)
843 \POS(0,0)\ar_{\vec n_{ij}}(0,-0.75)}
844 \POS(1.5,2.25)*{\bullet},*+!L{\vec u_0 = \vec w}
845 \POS(0,0)*{\bullet}
846 \POS(3,1.5)*+[o][F]{\scriptstyle 1}
847 \POS(0,2.5)*+[o][F]{\scriptstyle 2}
848 \POS(-3,1.5)*+[o][F]{\scriptstyle 3}
849 \POS(-3,-1.5)*+[o][F]{\scriptstyle 4}
850 \POS(0,-3)*+[o][F]{\scriptstyle 5}
851 \POS(3,-1.5)*+[o][F]{\scriptstyle 6}
852 \end{xy}
853 \end{minipage}
854 \begin{minipage}{0cm}
855 \begin{xy}
856 <\intercol,0pt>:<0pt,\intercol>::
857 \POS(-4,0)="a"\ar@[|(2)]@{-}(4,0)
858 \POS?(0)/\intercol/="b"\POS(2,0)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0i}}(0,0.75)}
859 \POS(2,0)*{\bullet},*+!U{\vec u_j}
860 \POS(-2,3)="a"\ar@[|(2)]@{-}(2,-3)
861 \POS?(0)/\intercol/="b"\POS(-1,1.5)*\xybox{"b"-"a":(0,0)\ar^{\vec n_{0j}}(0,+0.75)}
862 \POS(-1,1.5)*{\bullet},*+!R{\vec u_i}
863 \POS(-2,-3)="a"\ar@{-}(2,3)
864 \POS?(0)/\intercol/="b"\POS(-1.5,-2.25)
865 *\xybox{"b"-"a":(0,0)\ar^{\vec n_{ji}}(0,+0.75)
866 \POS(0,0)\ar_{\vec n_{ij}}(0,-0.75)}
867 \POS(-1.5,-2.25)*{\bullet},*+!L{\vec u_0 = \vec w}
868 \POS(0,0)*{\bullet}
869 \POS(3,1.5)*+[o][F]{\scriptstyle 1}
870 \POS(0,2.5)*+[o][F]{\scriptstyle 2}
871 \POS(-3,1.5)*+[o][F]{\scriptstyle 3}
872 \POS(-3,-1.5)*+[o][F]{\scriptstyle 4}
873 \POS(0,-3)*+[o][F]{\scriptstyle 5}
874 \POS(3,-1.5)*+[o][F]{\scriptstyle 6}
875 \end{xy}
876 \end{minipage}
877 \caption{Possible locations of $\vec w$ with respect to $\vec u_i$ and
878 $\vec u_j$, projected onto a plane orthogonal to the other rays, when
879 $\alpha_i \alpha_j > 0$.}
880 \label{fig:w:opposite}
881 \end{figure}
883 The algorithm is summarized in Algorithm~\ref{alg:closed}, where
884 we use the convention that in cone $K_i$, $\vec u_i$ refers to
885 $\vec u_0 = \vec w$.
886 Note that we do not need any of the rays or normals in this code.
887 The only information we need is the closedness of the facets in the
888 original cone and the signs of the $\alpha_i$.
890 \begin{algorithm}
891 \begin{tabbing}
892 next \= next \= next \= \kill
893 if $\alpha_j = 0$ \\
894 \> closed[$K_i$][$\vec u_j$] := closed[$\tilde K$][$\vec u_j$] \\
895 else if $i = j$ \\
896 \> if $\alpha_j > 0$ \\
897 \> \> closed[$K_i$][$\vec u_j$] := closed[$\tilde K$][$\vec u_j$] \\
898 \> else \\
899 \> \> closed[$K_i$][$\vec u_j$] := $\lnot$closed[$\tilde K$][$\vec u_j$] \\
900 else if $\alpha_i \alpha_j > 0$ \\
901 \> if closed[$\tilde K$][$\vec u_i$] = closed[$\tilde K$][$\vec u_j$] \\
902 \> \> closed[$K_i$][$\vec u_j$] := $i < j$ \\
903 \> else \\
904 \> \> closed[$K_i$][$\vec u_j$] := closed[$\tilde K$][$\vec u_j$] \\
905 else \\
906 \> closed[$K_i$][$\vec u_j$] := closed[$\tilde K$][$\vec u_i$] and
907 closed[$\tilde K$][$\vec u_j$]
908 \end{tabbing}
909 \caption{Determine whether the facet opposite $\vec u_j$ is closed in $K_i$.}
910 \label{alg:closed}
911 \end{algorithm}
913 \subsection{Triangulation in primal space}
914 \label{s:triangulation}
916 As in the case for Barvinok's decomposition (Section~\ref{s:decomposition}),
917 we can transform a triangulation of a (closed) cone into closed simple cones
918 into a triangulation of half-open simple cones that fully partitions the
919 original cone, i.e., such that the half-open simple cones do not intersect at their
920 facets.
921 Again, we apply Proposition~\ref{p:inclusion-exclusion} with $\vec y$
922 an interior point of the cone (Section~\ref{s:interior}).
923 Note that the interior point $\vec y$ may still intersect
924 some of the internal facets, so we may need to perturb it slightly.
925 In practice, we apply a lexicographical rule: for such (internal)
926 facets, which always appear in pairs, we close the one with
927 a lexico-positive normal and open the one with a lexico-negative
928 normal.
930 \subsection{Multivariate quasi-polynomials as lists of polynomials}
932 There are many definitions for a (univariate) \ai{quasi-polynomial}.
933 \shortciteN{Ehrhart1977} uses a definition based on {\em periodic number}s.
935 \begin{definition}
936 \label{d:periodic:1}
937 A rational \defindex{periodic number} $U(p)$
938 is a function $\ZZ \to \QQ$,
939 such that there exists a \defindex{period} $q$
940 such that $U(p) = U(p')$ whenever $p \equiv p' \mod q$.
941 \end{definition}
943 \begin{definition}
944 \label{d:qp:1}
945 A (univariate)
946 \defindex{quasi-polynomial}\/ $f$ of degree $d$ is
947 a function
949 f(n) = c_d(n) \, n^d + \cdots + c_1(n) \, n + c_0
952 where $c_i(n)$ are rational periodic numbers.
953 I.e., it is a polynomial expression of degree $d$
954 with rational periodic numbers for coefficients.
955 The \defindex{period} of a quasi-polynomial is the \ac{lcm}
956 of the periods of its coefficients.
957 \end{definition}
959 Other authors (e.g., \shortciteNP{Stanley1986})
960 use the following definition of a quasi-polynomial.
961 \begin{definition}
962 \label{d:qp:1:list}
963 A function $f : \ZZ \to \QQ$ is
964 a (univariate) \defindex{quasi-polynomial} of period $q$ if there
965 exists a list of $q$ polynomials $g_i \in \QQ[T]$ for $0 \le i < q$ such
966 that
968 f (s) = g_i(s) \qquad \hbox{if $s \equiv i \mod {q}$}
971 The functions $g_i$ are called the {\em constituents}.
972 \index{constituent}
973 \end{definition}
975 In our implementation, we use Definition~\ref{d:qp:1},
976 but whereas
977 \shortciteN{Ehrhart1977} uses a list of $q$ rational
978 numbers enclosed in square brackets to represent periodic
979 numbers, our periodic numbers are polynomial expressions
980 in fractional parts (Section~\ref{a:data}).
981 These fractional parts naturally extend to multivariate
982 quasi-polynomials.
983 The bracketed (``explicit'') periodic numbers can
984 be extended to multiple variables by nesting them
985 (e.g., \shortciteNP{Loechner1999}).
987 Definition~\ref{d:qp:1:list} could be extended in a similar way
988 by having a constituent for each residue modulo a vector period $\vec q$.
989 However, as pointed out by \citeN{Woods2006personal}, this may not result
990 in the minimum number of constituents.
991 A vector period can be considered as a lattice with orthogonal generators and
992 the number of constituents is equal to the index or determinant of that lattice.
993 By considering more general lattices, we can potentially reduce the number
994 of constituents.
995 \begin{definition}
996 \label{d:qp}
997 A function $f : \ZZ^n \to \QQ$ is
998 a (multivariate) \defindex{quasi-polynomial} of period $L$ if there
999 exists a list of $\det L$ polynomials $g_{\vec i} \in \QQ[T_1,\ldots,T_n]$
1000 for $\vec i$ in the fundamental parallelepiped of $L$ such
1001 that
1003 f (\vec s) = g_{\vec i}(\vec s) \qquad \hbox{if $\vec s \equiv \vec i \mod L$}
1006 \end{definition}
1008 To compute the period lattice from a fractional representation, we compute
1009 the appropriate lattice for each fractional part and then take their intersection.
1010 Recall that the argument of each fractional part is an affine expression
1011 in the parameters $(\sp a p + c)/m$,
1012 with $\vec a \in \ZZ^n$ and $c, m \in \ZZ$.
1013 Such a fractional part is translation invariant over
1014 any (integer) value of $\vec p$
1015 such that $\sp a p + m t = 0$, for some $\vec t \in \ZZ$.
1016 Solving this homogeneous equation over the integers (in our implementation,
1017 we use \PolyLib/'s \ai[\tt]{SolveDiophantine}) gives the general solution
1019 \begin{bmatrix}
1020 \vec p \\ t
1021 \end{bmatrix}
1023 \begin{bmatrix}
1024 U_1 \\ U_2
1025 \end{bmatrix}
1026 \vec x
1027 \qquad\text{for $\vec x \in \ZZ^n$}
1030 The matrix $U_1 \in \ZZ^{n \times n}$ then has the generators of
1031 the required lattice as columns.
1032 The constituents are computed by plugging in each integer point
1033 in the fundamental parallelepiped of the lattice.
1034 These points themselves are computed as explained in Section~\ref{s:fundamental}.
1035 Note that for computing the constituents, it is sufficient to take any
1036 representative of the residue class. For example, we could take
1037 $\vec w^\T = \vec k^\T W$ in the notations of Lemma~\ref{l:fundamental}.
1039 \begin{example}[\shortciteN{Woods2006personal}]
1040 Consider the parametric polytope
1042 P_{s,t}=\{\, x \mid 0 \le x \le (s+t)/2 \,\}
1045 The enumerator of $P_{s,t}$ is
1047 \begin{cases}
1048 \frac s 2 + \frac t 2 + 1 &
1049 \text{if $\begin{bmatrix}s \\ t \end{bmatrix} \in
1050 \begin{bmatrix}
1051 -1 & -2 \\ 1 & 0
1052 \end{bmatrix}
1053 \ZZ^2 +
1054 \begin{bmatrix}
1055 0 \\ 0
1056 \end{bmatrix}
1059 \frac s 2 + \frac t 2 + \frac 1 2 &
1060 \text{if $\begin{bmatrix}s \\ t \end{bmatrix} \in
1061 \begin{bmatrix}
1062 -1 & -2 \\ 1 & 0
1063 \end{bmatrix}
1064 \ZZ^2 +
1065 \begin{bmatrix}
1066 -1 \\ 0
1067 \end{bmatrix}
1070 \end{cases}
1072 The corresponding output of \ai[\tt]{barvinok\_enumerate} is
1073 \begin{verbatim}
1074 s + t >= 0
1075 1 >= 0
1077 Lattice:
1078 [[-1 1]
1079 [-2 0]
1081 [0 0]
1082 ( 1/2 * s + ( 1/2 * t + 1 )
1084 [-1 0]
1085 ( 1/2 * s + ( 1/2 * t + 1/2 )
1087 \end{verbatim}
1088 \end{example}
1090 \subsection{Left inverse of an affine embedding}
1091 \label{s:inverse}
1093 We often map a polytope onto a lower dimensional space to remove possible
1094 equalities in the polytope. These maps are typically represented
1095 by the inverse, mapping the coordinates $\vec x'$ of the lower-dimensional
1096 space to the coordinates $\vec x$ of (an affine subspace of) the original space,
1097 i.e.,
1099 \begin{bmatrix}
1100 \vec x \\ 1
1101 \end{bmatrix}
1103 \begin{bmatrix}
1104 T & \vec v \\ \vec 0^\T & 1
1105 \end{bmatrix}
1106 \begin{bmatrix}
1107 \vec x' \\ 1
1108 \end{bmatrix}
1111 where, as usual in \PolyLib/, we work with homogeneous coordinates.
1112 To obtain the transformation that maps the coordinates of the original
1113 space to the coordinates of the lower dimensional space,
1114 we need to compute the \ai{left inverse} of the above \ai{affine embedding},
1115 i.e., an $A$, $\vec b$ and $d$ such that
1118 \begin{bmatrix}
1119 \vec x' \\ 1
1120 \end{bmatrix}
1122 \begin{bmatrix}
1123 A & \vec b \\ \vec 0^\T & d
1124 \end{bmatrix}
1125 \begin{bmatrix}
1126 \vec x \\ 1
1127 \end{bmatrix}
1130 To compute this left inverse, we first compute the
1131 (right) \indac{HNF} of T,
1133 \begin{bmatrix}
1134 U_1 \\ U_2
1135 \end{bmatrix}
1138 \begin{bmatrix}
1139 H \\ 0
1140 \end{bmatrix}
1143 The left inverse is then simply
1145 \begin{bmatrix}
1146 d H^{-1}U_1 & -d H^{-1} \vec v \\ \vec 0^\T & d
1147 \end{bmatrix}
1150 We often also want a description of the affine subspace that is the range
1151 of the affine embedding and this is given by
1153 \begin{bmatrix}
1154 U_2 & - U_2 \vec v \\ \vec 0^T & 1
1155 \end{bmatrix}
1156 \begin{bmatrix}
1157 \vec x \\ 1
1158 \end{bmatrix}
1160 \vec 0
1163 This computation is implemented in \ai[\tt]{left\_inverse}.
1165 \subsection{Integral basis of the orthogonal complement of a linear subspace}
1166 \label{s:completion}
1168 Let $M_1 \in \ZZ^{m \times n}$ be a basis of a linear subspace.
1169 We first extend $M_1$ with zero rows to obtain a square matrix $M'$
1170 and then compute the (left) \indac{HNF} of $M'$,
1172 \begin{bmatrix}
1173 M_1 \\ 0
1174 \end{bmatrix}
1176 \begin{bmatrix}
1177 H & 0 \\ 0 & 0
1178 \end{bmatrix}
1179 \begin{bmatrix}
1180 Q_1 \\ Q_2
1181 \end{bmatrix}
1184 The rows of $Q_2$ span the orthogonal complement of the given subspace.
1185 Since $Q_2$ can be extended to a unimodular matrix, these rows form
1186 an integral basis.
1188 If the entries on the diagonal of $H$ are all $1$ then $M_1$
1189 can be extended to a unimodular matrix, by concatenating $M_1$ and $Q_2$.
1190 The resulting matrix is unimodular, since
1192 \begin{bmatrix}
1193 M_1 \\ Q_2
1194 \end{bmatrix}
1196 \begin{bmatrix}
1197 H & 0 \\ 0 & I_{n-m,n-m}
1198 \end{bmatrix}
1199 \begin{bmatrix}
1200 Q_1 \\ Q_2
1201 \end{bmatrix}
1204 This method for extending a matrix of which
1205 only a few lines are known to a \ai{unimodular matrix}
1206 is more general than the method described by \shortciteN{Bik1996PhD},
1207 which only considers extending a matrix given by a single row.
1209 \subsection{Ensuring a polyhedron has only revlex-positive rays}
1210 \label{s:revlexpos}
1212 The \ai[\tt]{barvinok\_series\_with\_options} function and all
1213 further \ai[\tt]{gen\_fun} manipulations assume that the effective
1214 parameter domain has only \ai{revlex-positive} rays.
1215 When used to computer \rgf/s, the \ai[\tt]{barvinok\_enumerate}
1216 application will therefore transform the effective parameter domain
1217 of a problem if it has revlex-negative rays.
1218 It will then not compute the generating function
1220 f(\vec x) = \sum_{\vec p \in \ZZ^m} \#(P_{\vec p} \cap \ZZ^d) \, x^{\vec p}
1225 g(\vec z) = \sum_{\vec p' \in \ZZ^n}
1226 \#(P_{T \vec p' + \vec t} \cap \ZZ^d) \, x^{\vec p'}
1228 instead, where $\vec p = T \vec p' + \vec t$,
1229 with $T \in \ZZ^{m \times n}$ and $\vec t \in \ZZ^m$, is an affine transformation
1230 that maps the transformed parameter space back to the original parameter space.
1232 First assume that the parameter domain does not contain any lines and
1233 that there are no equalities in the description of $P_{\vec p}$ that force
1234 the values of $\vec p$ for which $P_{\vec p}$ contains integer points
1235 to lie on a non-standard lattice.
1236 Let the effective parameter domain be given as
1238 \{\, \vec p \mid A \vec p + \vec c \ge \vec 0 \,\}
1240 where $A \in \ZZ^{s \times d}$ of row rank $d$;
1241 otherwise the effective parameter domain would contain a line.
1242 Let $H$ be the (left) \indac{HNF} of $A$, i.e.,
1244 A = H Q
1247 with $H$ lower-triangular with positive diagonal elements and
1248 $Q$ unimodular.
1249 Let $\tilde Q$ be the matrix obtained from $Q$ by reversing its rows,
1250 and, similarly, $\tilde H$ from $H$ by reversing the columns.
1251 After performing the transformation
1252 $\vec p' = \tilde Q \vec p$, i.e.,
1253 $\vec p = \tilde Q^{-1} \vec p'$, the transformed parameter domain
1254 is given by
1256 \{\, \vec p' \mid A \tilde Q^{-1} \vec p' + \vec c \ge \vec 0 \,\}
1260 \{\, \vec p' \mid \tilde H \vec p' + \vec c \ge \vec 0 \,\}
1263 The first constraint of this domain is
1264 $h_{11} p'_m + c_1 \ge 0$. A ray with non-zero final coordinate
1265 therefore has a positive final coordinate.
1266 Similarly, the second constraint is
1267 $h_{22} p'_{m-1} + h_{21} p'_m + c_2 \ge 0$.
1268 A ray with zero $n$th coordinate, but non-zero $n-1$st coordinate,
1269 will therefore have a positive $n-1$st coordinate.
1270 Continuing this reasoning, we see that all rays in the transformed
1271 domain are revlex-positive.
1273 If the parameter domain does contains lines, but is not restricted
1274 to a non-standard lattice, then the number of points in the parametric
1275 polytope is invariant over a translation along the lines.
1276 It is therefore sufficient to compute the number of points in the
1277 orthogonal complement of the linear subspace spanned by the lines.
1278 That is, we apply a prior transformation that maps a reduced parameter
1279 domain to this subspace,
1281 \vec p = L^\perp \vec p' =
1282 \begin{bmatrix}
1283 L & L^\perp
1284 \end{bmatrix}
1285 \begin{bmatrix}
1286 0 \\ I
1287 \end{bmatrix}
1288 \vec p'
1291 where $L$ has the lines as columns, and $L^\perp$ an integral basis
1292 for the orthogonal complement (Section~\ref{s:completion}).
1293 Note that the inverse transformation
1295 \vec p' =
1296 L^{-\perp}
1297 \vec p =
1298 \begin{bmatrix}
1299 0 & I
1300 \end{bmatrix}
1301 \begin{bmatrix}
1302 L & L^\perp
1303 \end{bmatrix}^{-1}
1304 \vec p
1306 has integral coefficients since $L^\perp$ can be extended to a unimodular matrix.
1308 If the parameter values $\vec p$ for which $P_{\vec p}$ contains integer points
1309 are restricted to a non-standard lattice, we first replace the parameters
1310 by a different set of parameters that lie on the standard lattice
1311 through ``\ai{parameter compression}''\shortcite{Meister2004PhD},
1313 \vec p = C \vec p'
1316 The (left) inverse of $C$ can be computes as explained in
1317 Section~\ref{s:inverse}, giving
1319 \vec p' = C^{-L} \vec p
1322 We have to be careful to only apply this transformation when
1323 both the equalities computed in Section~\ref{s:inverse} are satisfied
1324 and some additional divisibility constraints.
1325 In particular if $\vec a^\T/d$ is a row of $C^{-L}$, with $\vec a \in \ZZ^{n'}$
1326 and $d \in \ZZ$, the transformation can only be applied to parameter values
1327 $\vec p$ such that $d$ divides $\sp a p$.
1329 The complete transformation is given by
1331 \vec p = C L^\perp \hat Q^{-1} \vec p'
1335 \vec p' = \hat Q L^{-\perp} C^{-L} \vec p
1339 \subsection{Parametric Volume Computation}
1341 The \ai{volume} of a (parametric) polytope can serve as an approximation
1342 for the number of integer points in the polytope.
1343 We basically follow the description of~\shortciteN{Rabl2006} here, except that we
1344 focus on volume computation for {\em linearly}
1345 parametrized polytopes, which we exploit to determine the sign
1346 of the determinants we compute, as explained below.
1348 Note first that
1349 the vertices of a linearly parametrized polytope are affine expressions
1350 in the parameters that may be valid only in parts (chambers)
1351 of the parameter domain.
1352 Since the volume computation is based on the (active) vertices, we perform
1353 the computation in each chamber separately.
1354 Also note that since the vertices are affine expressions, it is
1355 easy to check whether they belong to a facet.
1357 The volume of a $d$-simplex, i.e., a $d$-dimensional polytope with
1358 $d+1$ vertices, is relatively easy to compute.
1359 In particular, if $\vec v_i(\vec p)$, for $0 \le i \le d$,
1360 are the (parametric) vertices
1361 of the simplex $P$ then
1362 \begin{equation}
1363 \label{eq:volume}
1364 \vol P =
1365 \frac 1 {d!}
1366 \left|
1367 \det
1368 \begin{bmatrix}
1369 v_{11}(\vec p) - v_{01}(\vec p) &
1370 v_{12}(\vec p) - v_{02}(\vec p) &
1371 \ldots &
1372 v_{1d}(\vec p) - v_{0d}(\vec p)
1374 v_{21}(\vec p) - v_{01}(\vec p) &
1375 v_{22}(\vec p) - v_{02}(\vec p) &
1376 \ldots &
1377 v_{2d}(\vec p) - v_{0d}(\vec p)
1379 \vdots & \vdots & \ddots & \vdots
1381 v_{d1}(\vec p) - v_{01}(\vec p) &
1382 v_{d2}(\vec p) - v_{02}(\vec p) &
1383 \ldots &
1384 v_{dd}(\vec p) - v_{0d}(\vec p)
1385 \end{bmatrix}
1386 \right|.
1387 \end{equation}
1389 If $P$ is not a simplex, i.e., $N > d+1$, with $N$ the number of
1390 vertices of $P$, then the standard way of computing the volume of $P$
1391 is to first {\em triangulate} $P$, i.e., subdivide $P$ into simplices,
1392 and then to compute and sum the volumes of the resulting simplices.
1393 One way of computing a triangulation is to
1394 compute the \ai{barycenter}
1396 \frac 1 N \sum_i \vec v_i(\vec p)
1398 of $P$
1399 and to perform a subdivision by computing the convex hulls
1400 of the barycenter with each of the facets of $P$.
1401 If a given facet of $P$ is itself a simplex, then this convex hull
1402 is also a simplex. Otherwise the facet is further subdivided.
1403 This recursive process terminates as every $1$-dimensional polytope
1404 is a simplex.
1406 The triangulation described above is known as
1407 the boundary triangulation~\shortcite{Bueler2000exact} and is used
1408 by \shortciteN{Rabl2006} in his implementation.
1409 The Cohen-Hickey triangulation~\shortcite{Cohen1979volumes,Bueler2000exact}
1410 is a much more efficient variation and uses one of the vertices
1411 instead of the barycenter. The facets incident on the vertex
1412 do not have to be considered in this case because the resulting subpolytopes
1413 would have zero volume.
1414 Another possibility is to use a
1415 ``lifting'' triangulation~\shortcite{Lee1991,DeLoera1995}.
1416 In this triangulation, each vertex is assigned a (random) ``height'' in
1417 an extra dimension.
1418 The projection of the ``lower envelope'' of the resulting polytope onto
1419 the original space results in a subdivision, which is a triangulation
1420 with very high probability.
1422 A complication with the lifting triangulation is that the constraint system
1423 of the lifted polytope will in general not be linearly parameterized,
1424 even if the original polytope is.
1425 It is, however, sufficient to perform the triangulation for a particular
1426 value of the parameters inside the chamber since the parametric polytope
1427 has the same combinatorial structure throughout the chamber.
1428 The triangulation obtained for the instantiated vertices can then
1429 be carried over to the corresponding parametric vertices.
1430 We only need to be careful to select a value for the parameters that
1431 does not lie on any facet of the chambers. On these chambers, some
1432 of the vertices may coincide.
1433 For linearly parametrized polytopes, it is easy to find a parameter
1434 point in the interior of a chamber, as explained in Section~\ref{s:interior}.
1435 Note that this point need not be integer.
1437 A direct application of the above algorithm, using any of the triangulations,
1438 would yield for each chamber
1439 a volume expressed as the sum of the absolute values of polynomials in
1440 the parameters. To remove the absolute value, we plug in a particular
1441 value of the parameters (not necessarily integer)
1442 belonging to the given chamber for which we know that the volume is non-zero.
1443 Again, it is sufficient to take any point in the interior of the chamber.
1444 The sign of the resulting value then determines the sign of the whole
1445 polynomial since polynomials are continuous functions and will not change
1446 sign without passing through zero.
1448 \subsection{Maclaurin series division}
1449 \label{s:division}
1451 If $P(t)$ and $Q(t)$ are two Maclaurin series
1452 \begin{align*}
1453 P(t) & = a_0 + a_1 t + a_2 t^2 + \cdots \\
1454 Q(t) & = b_0 + b_1 t + b_2 t^2 + \cdots
1456 \end{align*}
1457 then, as outlined by \shortciteN[241--247]{Henrici1974},
1458 we can compute the coefficients $c_l$ in
1460 \frac{P(t)}{Q(t)} =: c_0 + c_1 t + c_2 t^2 + \cdots
1462 by applying the recurrence relation
1464 c_l = \frac 1 {b_0} \left( a_l - \sum_{i=1}^l b_i c_{l-i} \right)
1467 To avoid dealing with denominators, we can also compute
1468 $d_l = b_0^{l+1} c_l$ instead as
1470 d_l = b_0^l a_l - \sum_{i=1}^l b_0^{i-1} b_i c_{l-i}
1473 The coefficients $c_l$ can then be directly read off as
1475 c_l = \frac{d_l}{b_0^{l+1}}
1479 \subsection{Specialization through exponential substitution}
1481 This section draws heavily from \shortciteN{Koeppe2006experiments}.
1483 We define a ``short'' \defindex{\rgf/} to be a function of the form
1484 \begin{equation}
1485 \label{eq:rgf}
1486 f(\vec x)=
1487 \sum_{i\in I}\alpha_i
1488 \frac{\sum_{k=1}^{r} \vec x^{\vec w_{ik} }}
1489 {\prod_{j=1}^{k_i}\left(1-\vec x^{\vec b_{ij}}\right)}
1491 \end{equation}
1492 with $\vec x \in \CC^d$, $\alpha_i \in \QQ$,
1493 $\vec w_{i k} \in \ZZ^d$ and $\vec b_{i j} \in \ZZ^d \setminus \{\vec 0\}$.
1495 After computing the \rgf/~\eqref{eq:rgf} of a polytope
1496 (with $k_i = d$ for all $i$),
1497 the number of lattice points in the polytope can be obtained
1498 by evaluating $f(\vec 1)$. Since $\vec 1$ is a pole of each
1499 term, we need to compute the constant term in the Laurent expansions
1500 of each term in~\eqref{eq:rgf} about $\vec 1$.
1501 Since it is easier to work with univariate series, a substitution is usually
1502 applied, either a \ai{polynomial substitution}
1504 \vec x = (1+t)^{\vec \lambda}
1507 as implemented in \LattE/ \shortcite{latte1.1},
1508 or an \ai{exponential substitution} (see, e.g., \shortciteNP{Barvinok1999}),
1510 \vec x = e^{t \vec \lambda}
1513 as implemented in \LattEmk/ \shortcite{latte-macchiato}.
1514 In each case, $\vec \lambda \in \ZZ^d$ is a vector that is not orthogonal
1515 to any of the $\vec b_{ij}$.
1516 Both substitutions also transform the problem of computing the
1517 constant term in the Laurent expansions about $\vec x = \vec 1$
1518 to that of computing the constant term in the
1519 Laurent expansions about $t = 0$.
1521 Consider now one of the terms in~\eqref{eq:rgf},
1523 g(t) =
1524 \frac{\sum_{k=1}^{r} e^{a_k t}}
1525 {\prod_{j=1}^{d}\left(1-e^{c_j t}\right)}
1528 with $a_k = \sp{w_{ik}}{\lambda}$ and $c_j = \sp{b_{ij}}{\lambda}$.
1529 We rewrite this equation as
1531 g(t) =
1532 (-1)^d
1533 \frac{\sum_{k=1}^{r} e^{a_k t}}
1534 {t^d \prod_{j=1}^d c_j}
1535 \prod_{j=1}^d \frac{-c_j t}
1536 {1-e^{c_j t}}
1539 The second factor is analytic in a neighborhood of the origin
1540 $x = c_1 = \cdots = c_d = 0$ and therefore has a Taylor series expansion
1541 \begin{equation}
1542 \label{eq:todd}
1543 \prod_{j=1}^d \frac{-c_j t}
1544 {1-e^{c_j t}}
1546 \sum_{m=0}^{\infty} \todd_m(-c_1, \ldots, -c_d) t^m
1548 \end{equation}
1549 where $\todd_m$ is a homogeneous polynomial of degree $m$ called
1550 the $m$-th \ai{Todd polynomial}~\cite{Barvinok1999}.
1551 Also expanding the numerator in the first factor, we find
1553 g(t) = \frac{(-1)^d}{t^d \prod_{j=1}^d c_j}
1554 \left(
1555 \sum_{n=0}^{\infty}\frac{\sum_{k=1}^{r} a_k^n}{n!}
1556 \right)
1557 \left(
1558 \sum_{m=0}^{\infty} \todd_m(-c_1, \ldots, -c_d) t^m
1559 \right)
1562 with constant term
1563 \begin{multline}
1564 \label{eq:todd:constant}
1565 \frac{(-1)^d}{t^d \prod_{j=1}^d c_j}
1566 \left(\sum_{i=0}^d \frac{\sum_{k=1}^{r} a_k^i}{i!}
1567 \todd_{d-i}(-c_1, \ldots, -c_d)\right)t^d
1568 = \\
1569 \frac{(-1)^d}{\prod_{j=1}^d c_j}
1570 \sum_{i=0}^d \frac{\sum_{k=1}^{r} a_k^i}{i!} \todd_{d-i}(-c_1, \ldots, -c_d)
1572 \end{multline}
1573 To compute the first $d+1$ terms in the Taylor series~\eqref{eq:todd},
1574 we write down the truncated Taylor series
1576 \frac{e^t -1}t \equiv
1577 \sum_{i=0}^d \frac 1{(i+1)!} t^i \equiv
1578 \frac 1 {(d+1)!} \sum_{i=0}^d \frac{(d+1)!}{(i+1)!} t^i
1579 \mod t^{d+1}
1582 where we have
1584 \frac 1 {(d+1)!} \sum_{i=0}^d \frac{(d+1)!}{(i+1)!} t^i
1585 \in \frac 1{(d+1)!} \ZZ[t]
1588 Computing the reciprocal as explained in Section~\ref{s:division},
1589 we find
1590 \begin{equation}
1591 \label{eq:t-exp-1}
1592 \frac{t}{e^t-1} = \frac 1{\frac{e^t -1}t}
1593 \equiv (d+1)! \frac 1{\sum_{i=0}^d \frac{(d+1)!}{(i+1)!} t^i}
1594 =: \sum_{i=0}^d b_i t^i
1596 \end{equation}
1597 Note that the constant term of the denominator is $1/(d+1)!$.
1598 The denominators of the quotient are therefore $((d+1)!)^{i+1}/(d+1)!$.
1599 Also note that the $b_i$ are independent of the generating function
1600 and can be computed in advance.
1601 An alternative way of computing the $b_i$ is to note that
1603 \frac{t}{e^t-1} = \sum_{i=0}^\infty B_i \frac{t^i}{i!}
1606 with $B_i = i! \, b_i$ the \ai{Bernoulli number}s, which can be computed
1607 using the recurrence~\eqref{eq:Bernoulli} (see Section~\ref{s:nested}).
1609 Substituting $t$ by $c_j t$ in~\eqref{eq:t-exp-1}, we have
1611 \frac{-c_j t}{1-e^{c_j t}} = \sum_{i=0}^d b_i c_j^i t^i
1614 Multiplication of these truncated Taylor series for each $c_j$
1615 results in the first $d+1$ terms of~\eqref{eq:todd},
1617 \sum_{m=0}^{d} \todd_m(-c_1, \ldots, -c_d) t^m
1619 \sum_{m=0}^{d} \frac{\beta_m}{((d+1)!)^m} t^m
1622 from which
1623 it is easy to compute the constant term~\eqref{eq:todd:constant}.
1624 Note that this convolution can also be computed without the use
1625 of rational coefficients,
1627 \frac{(-1)^d}{\prod_{j=1}^d c_j}
1628 \sum_{i=0}^d \frac{\alpha_i}{i!} \frac{\beta_{d-i}}{((d+1)!)^{d-i}}
1630 \frac{(-1)^d}{((d+1)!)^d\prod_{j=1}^d c_j}
1631 \sum_{i=0}^d \left(\frac{((d+1)!)^i}{i!}\alpha_i\right) \beta_{d-i}
1634 with $\alpha_i = \sum_{k=1}^{r} a_k^i$.
1636 \begin{example}
1637 Consider the \rgf/
1638 \begin{multline*}
1639 \f T x =
1640 \frac{x_1^2}{(1-x_1^{-1})(1-x_1^{-1}x_2)}
1642 \frac{x_2^2}{(1-x_2^{-1})(1-x_1 x_2^{-1})}
1643 + {} \\
1644 \frac1{(1-x_1)(1-x_2)}
1645 \end{multline*}
1646 from \shortciteN[Example~39]{Verdoolaege2005PhD}.
1647 Since this is a 2-dimensional problem, we first compute the first
1648 3 Todd polynomials (evaluated at $-1$),
1650 \frac{e^t -1}t \equiv
1651 1 + \frac 1 2 t + \frac 1 7 t^2 =
1652 \frac 1 6
1653 \begin{bmatrix}
1654 6 & 3 & 1
1655 \end{bmatrix}
1659 \frac t{e^t -1} \equiv
1660 \begin{bmatrix}
1661 \displaystyle\frac{-1}{1} & \displaystyle\frac{3}{6} & \displaystyle\frac{-3}{36}
1662 \end{bmatrix}
1665 where we represent each truncated power series by a vector of its
1666 coefficients.
1667 The vector $\vec\lambda = (1, -1)$ is not
1668 orthogonal to any of the rays, so we can use the substitution
1669 $\vec x = e^{(1, -1)t}$
1670 and obtain
1672 \frac{e^{2t}}{(1-e^{-t})(1-e^{-2t})}
1674 \frac{e^{-2t}}{(1-e^{t})(1-e^{2t})}
1676 \frac1{(1-e^{t})(1-e^{-t})}
1679 We have
1680 \begin{align*}
1681 \frac{t}{1-e^{- t}} & =
1682 \begin{bmatrix}
1683 \displaystyle\frac{-1}{1} & \displaystyle\frac{-3}{6} & \displaystyle\frac{-3}{36}
1684 \end{bmatrix}
1686 \frac{2t}{1-e^{-2 t}} & =
1687 \begin{bmatrix}
1688 \displaystyle\frac{-1}{1} & \displaystyle\frac{-6}{6} & \displaystyle\frac{-12}{36}
1689 \end{bmatrix}
1691 \frac{-t}{1-e^{t}} & =
1692 \begin{bmatrix}
1693 \displaystyle\frac{-1}{1} & \displaystyle\frac{3}{6} & \displaystyle\frac{-3}{36}
1694 \end{bmatrix}
1696 \frac{-2t}{1-e^{2t}} & =
1697 \begin{bmatrix}
1698 \displaystyle\frac{-1}{1} & \displaystyle\frac{6}{6} & \displaystyle\frac{-12}{36}
1699 \end{bmatrix}
1701 \end{align*}
1702 The first term in the \rgf/ evaluates to
1703 \begin{align*}
1705 \frac 1{-1 \cdot -2}
1706 \begin{bmatrix}
1707 \displaystyle\frac{1}{1} & \displaystyle\frac{2}{1} & \displaystyle\frac{4}{2}
1708 \end{bmatrix}
1710 \left(
1711 \begin{bmatrix}
1712 \displaystyle\frac{-1}{1} & \displaystyle\frac{-3}{6} & \displaystyle\frac{-3}{36}
1713 \end{bmatrix}
1714 \begin{bmatrix}
1715 \displaystyle\frac{-1}{1} & \displaystyle\frac{-6}{6} & \displaystyle\frac{-12}{36}
1716 \end{bmatrix}
1717 \right)
1719 = {} &
1720 \frac 1{2}
1721 \begin{bmatrix}
1722 \displaystyle\frac{1}{1} & \displaystyle\frac{2}{1} & \displaystyle\frac{4}{2}
1723 \end{bmatrix}
1725 \begin{bmatrix}
1726 \displaystyle\frac{1}{1} & \displaystyle\frac{9}{6} & \displaystyle\frac{33}{36}
1727 \end{bmatrix}
1729 = {} &
1730 \frac 1{72}
1731 \begin{bmatrix}
1732 1 & 2 \cdot 6 & 4 \cdot 18
1733 \end{bmatrix}
1735 \begin{bmatrix}
1736 1 & 9 & 33
1737 \end{bmatrix}
1738 = \frac {213}{72} = \frac{71}{24}
1740 \end{align*}
1741 Due to symmetry, the second term evaluates to the same value,
1742 while for the third term we find
1744 \frac{1}{-1\cdot 1 \cdot 36}
1745 \begin{bmatrix}
1746 1 & 0 \cdot 6 & 0 \cdot 18
1747 \end{bmatrix}
1749 \begin{bmatrix}
1750 1 & 0 & -3
1751 \end{bmatrix}
1753 \frac{-3}{-36} = \frac 1{12}
1756 The sum is
1758 \frac{71}{24} + \frac{71}{24} + \frac 1{12} = 6
1761 \end{example}
1763 Note that the run-time complexities of polynomial and exponential
1764 substitution are basically the same. The experiments of
1765 \citeN{Koeppe2006primal} are somewhat misleading in this respect
1766 since the polynomial substitution (unlike the exponential
1767 substitution) had not been optimized to take full
1768 advantage of the stopped Barvinok decomposition.
1769 For comparison, \autoref{t:hickerson} shows running times
1770 for the same experiments of that paper, but using
1771 barvinok version \verb+barvinok-0.23-47-gaa9024e+
1772 on an Athlon MP 1500+ with 512MiB internal memory.
1773 This machine appears to be slightly slower than the
1774 machine used in the experiments of \citeN{Koeppe2006primal}
1775 as computing {\tt hickerson-14} using the dual decomposition
1776 with polynomial substitution an maximal index 1
1777 took 2768 seconds on this machine using \LattEmk/.
1778 At this stage, it is not clear yet why the number of
1779 cones in the dual decomposition of {\tt hickerson-13}
1780 differs from that of \LattE/~\shortcite{latte1.1} and
1781 \LattEmk/~\cite{latte-macchiato}.
1782 We conclude from \autoref{t:hickerson} that (our implementation of)
1783 the exponential substitution is always slightly faster than
1784 (our implementation of) the polynomial substitution.
1785 The optimal maximal index for these examples is about 500,
1786 which agrees with the experiments of \citeN{Koeppe2006primal}.
1788 \begin{table}
1789 \begin{center}
1790 \begin{tabular}{rrrrrrr}
1791 \hline
1793 \multicolumn{3}{c}{Dual decomposition} &
1794 \multicolumn{3}{c}{Primal decomposition}
1797 & \multicolumn{2}{c}{Time (s)} &
1798 & \multicolumn{2}{c}{Time (s)}
1800 \cline{3-4}
1801 \cline{6-7}
1802 Max.\ index & Cones & Poly & Exp & Cones & Poly & Exp \\
1803 \hline
1804 \multicolumn{7}{c}{{\tt hickerson-12}}
1806 \hline
1807 1 & 11625 & 9.24 & 8.90 & 7929 & 4.80 & 4.55
1809 10 & 4251 & 4.32 & 4.19 & 803 & 0.66 & 0.62
1811 100 & 980 & 1.42 & 1.35 & 84 & 0.13 & 0.12
1813 200 & 550 & 1.00 & 0.92 & 76 & 0.12 & 0.12
1815 300 & 474 & 0.93 & 0.86 & 58 & 0.12 & 0.10
1817 500 & 410 & 0.90 & 0.83 & 42 & 0.10 & 0.10
1819 1000 & 130 & 0.42 & 0.38 & 22 & {\bf 0.10} & {\bf 0.07}
1821 2000 & 10 & {\bf 0.10} & {\bf 0.10} & 22 & 0.10 & 0.09
1823 5000 & 7 & 0.12 & 0.11 & 7 & 0.12 & 0.10
1825 \hline
1826 \multicolumn{7}{c}{{\tt hickerson-13}}
1828 \hline
1829 1 & 494836 & 489 & 463 & 483507 & 339 & 315
1831 10 & 296151 & 325 & 309 & 55643 & 51 & 48
1833 100 & 158929 & 203 & 192 & 9158 & 11 & 10
1835 200 & 138296 & 184 & 173 & 6150 & 9 & 8
1837 300 & 110438 & 168 & 157 & 4674 & 8 & 7
1839 500 & 102403 & 163 & 151 & 3381 & {\bf 8} & {\bf 7}
1841 1000 & 83421 & {\bf 163} & {\bf 149} & 2490 & 8 & 7
1843 2000 & 77055 & 170 & 153 & 1857 & 10 & 8
1845 5000 & 57265 & 246 & 211 & 1488 & 13 & 11
1847 10000 & 50963 & 319 & 269 & 1011 & 26 & 21
1849 \hline
1850 \multicolumn{7}{c}{{\tt hickerson-14}}
1852 \hline
1853 1 & 1682743 & 2171 & 2064 & 552065 & 508 & 475
1855 10 & 1027619 & 1453 & 1385 & 49632 & 62 & 59
1857 100 & 455474 & 768 & 730 & 8470 & 14 & 13
1859 200 & 406491 & 699 & 661 & 5554 & 11 & 10
1861 300 & 328340 & 627 & 590 & 4332 & 11 & 9
1863 500 & 303566 & 605 & 565 & 3464 & {\bf 11} & {\bf 9}
1865 1000 & 232626 & {\bf 581} & {\bf 532} & 2384 & 12 & 10
1867 2000 & 195368 & 607 & 545 & 1792 & 14 & 12
1869 5000 & 147496 & 785 & 682 & 1276 & 19 & 16
1871 10000 & 128372 & 966 & 824 & 956 & 29 & 23
1873 \hline
1874 \end{tabular}
1875 \caption{Timing results of dual and primal decomposition with
1876 polynomial or exponential substitution on the Hickerson examples}
1877 \label{t:hickerson}
1878 \end{center}
1879 \end{table}
1881 \subsection{Approximate Enumeration using Nested Sums}
1882 \label{s:nested}
1884 If $P \in \QQ^d$ is a polyhedron and $p(\vec x) \in \QQ[\vec x]$ is a
1885 polynomial and we want to sum $p(\vec x)$ over all integer values
1886 of (a subset of) the variables $\vec x$, then we can do this incrementally
1887 by taking a variable $x_1$ with lower bound $L(\vec{\hat x})$
1888 and upper bound $U(\vec{\hat x})$, with $\vec{\hat x} = (x_2, \ldots, x_d)$,
1889 and computing
1890 \begin{equation}
1891 \label{eq:nested:sum}
1892 Q(\vec{\hat x}) = \sum_{x_1 = L(\vec{\hat x})}^{U(\vec{\hat x})} p(\vec x)
1894 \end{equation}
1895 Since $P$ is a polytope, the lower bound is a maximum of affine expressions
1896 in the remaining variables, while the upper bound is a minimum of such expressions.
1897 If the coefficients in these expressions are all integer, then we can
1898 compute $Q(\vec{\hat x})$ exactly as a piecewise polynomial using formulas
1899 for sums of powers, as proposed by, e.g.,
1900 \shortciteN{Tawbi1994,Sakellariou1997sums,VanEngelen2004}.
1901 If some of the coefficients are not integer, we can apply the same formulas
1902 to obtain an approximation, which can is some cases be shown
1903 to be an overapproximation~\shortcite{VanEngelen2004}.
1904 Note that if we take the initial polynomial to be the constant $1$, then
1905 this gives us a method for computing an approximation of the number
1906 of integer points in a (parametric) polytope.
1908 The first step is to compute the chamber decomposition of $P$ when viewed
1909 as a 1-dimensional parametric polytope. That is, we need to partition
1910 the projection of $P$ onto the remaining variables into polyhedral cells
1911 such that in each cell, both the upper and the lower bound are described
1912 by a single affine expression. Basically, for each pair of lower and upper
1913 bound, we compute the cell where the chosen lower bound is (strictly)
1914 smaller than all other lower bounds and similarly for the upper bound.
1916 For any given pair of lower and upper bound $(l(\vec {\hat x}), u(\vec{\hat x}))$,
1917 the formula~\eqref{eq:nested:sum} is computed for each monomial of $p(\vec x)$
1918 separately. For the constant term $\alpha_0$, we have
1919 \begin{equation}
1920 \label{eq:summation:1d}
1921 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_0(\vec{\hat x})
1922 = \alpha_0(\vec{\hat x}) \left(u(\vec{\hat x}) - l(\vec {\hat x}) + 1\right)
1924 \end{equation}
1925 For the higher degree monomials, we use the formula
1926 \begin{equation}
1927 \label{eq:summation}
1928 \sum_{k=0}^{m-1} k^n = {1\over{n+1}}\sum_{k=0}^n{n+1\choose{k}} B_k m^{n+1-k}
1929 =: S_n(m)
1931 \end{equation}
1932 with $B_i$ the \ai{Bernoulli number}s, which can be computed
1933 using the recurrence
1934 \begin{equation}
1935 \label{eq:Bernoulli}
1936 \sum_{j=0}^m{m+1\choose{j}}B_j = 0
1937 \qquad B_0 = 1
1939 \end{equation}
1940 Note that \eqref{eq:summation} is also valid if $m = 0$,
1941 i.e., $S_n(0) = 0$, a fact
1942 that can be easily shown using Newton series~\shortcite{VanEngelen2004}.
1944 \newcounter{saveenumi}
1946 Since we can only directly apply the summation formula when
1947 the lower bound is zero (or one), we need to consider several
1948 cases.
1949 \begin{enumerate}
1950 \item $l(\vec {\hat x}) \ge 1$
1951 \label{i:l}
1953 \begin{align*}
1954 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1956 \alpha_n(\vec{\hat x})
1957 \left(
1958 \sum_{x_1 = 1}^{u(\vec{\hat x})} x_1^n
1960 \sum_{x_1 = 1}^{l(\vec {\hat x})-1} x_1^n
1961 \right)
1964 \alpha_n(\vec{\hat x})
1965 \left( S_n(u(\vec{\hat x})+1) - S_n(l(\vec {\hat x})) \right)
1966 \end{align*}
1968 \item $u(\vec{\hat x}) \le -1$
1969 \label{i:u}
1971 \begin{align*}
1972 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1974 \alpha_n(\vec{\hat x}) (-1)^n
1975 \sum_{x_1 = -u(\vec {\hat x})}^{-l(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1978 \alpha_n(\vec{\hat x}) (-1)^n
1979 \left( S_n(-l(\vec{\hat x})+1) - S_n(-u(\vec {\hat x})) \right)
1980 \end{align*}
1982 \item $l(\vec {\hat x}) \le 0$ and $u(\vec{\hat x}) \ge 0$
1983 \label{i:l:u}
1985 \begin{align*}
1986 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1988 \alpha_n(\vec{\hat x})
1989 \left(
1990 \sum_{x_1 = 0}^{u(\vec{\hat x})} x_1^n
1992 (-1)^n
1993 \sum_{x_1 = 1}^{-l(\vec {\hat x})} x_1^n
1994 \right)
1997 \alpha_n(\vec{\hat x})
1998 \left(
1999 S_n(u(\vec{\hat x})+1)
2001 (-1)^n
2002 S_n(-l(\vec{\hat x})+1)
2003 \right)
2004 \end{align*}
2006 \setcounter{saveenumi}{\value{enumi}}
2007 \end{enumerate}
2009 If the coefficients in the lower and upper bound are all
2010 integer, then the above 3 cases partition (the integer points in)
2011 the projection of $P$ onto the remaining variables.
2012 However, if some of the coefficients are rational, then the lower
2013 and upper bound can lie in the open interval $(0,1)$ for some
2014 values of $\vec{\hat x}$. We may therefore also want to consider
2015 the following two cases.
2017 \begin{enumerate}
2018 \setcounter{enumi}{\value{saveenumi}}
2019 \item $0 < l(\vec {\hat x}) < 1$
2020 \label{i:l:fractional}
2023 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
2025 \alpha_n(\vec{\hat x})
2026 S_n(u(\vec{\hat x})+1)
2029 \item $0 < -u(\vec {\hat x}) < 1$
2030 \label{i:u:fractional}
2033 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
2035 \alpha_n(\vec{\hat x})
2036 (-1)^n
2037 S_n(-l(\vec{\hat x})+1)
2040 \end{enumerate}
2041 Note that we may add the constraint $u \ge 1$ to
2042 case~\ref{i:l:fractional} and the constraint $l \le -1$
2043 to case~\ref{i:u:fractional}, since the correct value for
2044 these two cases would be zero if these extra constraints do not hold.
2046 An alternative to adding the above two cases would be
2047 to simply ignore them, i.e., assume a value of $0$.
2048 Another alternative would be to reduce case~\ref{i:l:u}
2051 l(\vec {\hat x}) \le -1\quad\hbox{and}\quad u(\vec{\hat x}) \ge 1
2054 while extending cases~\ref{i:l:fractional} and~\ref{i:u:fractional}
2057 -1 < l(\vec {\hat x}) < 1\quad\hbox{and}\quad u \ge 1
2061 -1 < u(\vec {\hat x}) < 1\quad\hbox{and}\quad l \le -1
2064 respectively, with the remaining cases
2065 ($-1 < l \le u < 1$) having value $0$.
2066 There does not appear to be a consistently better choice
2067 here, as each of these three approaches seems to yield better
2068 results on some examples.
2069 The last approach has the additional drawback that we
2070 would also have to deal with 5 cases, even if the bounds
2071 are integers.
2073 If at least one of the lower or upper bound is an
2074 integer affine expression, then we can reduce
2075 the 3 (or 5) cases to a single case (case~\ref{i:l:u})
2076 by an affine substitution that ensure that the
2077 new (lower or upper) bound is zero.
2078 In particular, if $l(\vec {\hat x})$ is an integer affine
2079 expression, then we replace $x$ by $x' + l(\vec {\hat x})$
2080 and similarly for an upper bound.
2082 \subsection{Exact Enumeration using Nested Sums}
2083 \label{s:nested:exact}
2085 The exact enumeration using nested sums proceeds in much
2086 the same way as the approximate enumeration from
2087 \autoref{s:nested}, with the notable exception
2088 that we need to take the (greatest or least) integer part
2089 of any fractional bounds that may occur.
2090 This has several consequences, discussed below.
2092 Since we will introduce floors during the recursive application
2093 of the procedure, we may as well allow the weight
2094 $p(\vec x)$ in~\eqref{eq:nested:sum} to be a (piecewise)
2095 quasipolynomial.
2097 For the constant term, \eqref{eq:summation:1d} becomes
2099 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_0(\vec{\hat x})
2100 = \alpha_0(\vec{\hat x})
2101 \left(\floor{u(\vec{\hat x})} - \ceil{l(\vec {\hat x})} + 1\right)
2105 Since we force the lower and upper bounds to be integers,
2106 cases~\ref{i:l:fractional} and~\ref{i:u:fractional} do not occur,
2107 while the conditions for cases~\ref{i:l} and~\ref{i:u} can be simplified
2110 l(\vec {\hat x}) > 0
2114 u(\vec {\hat x}) < 0
2117 respectively.
2119 If the variable $x$ appears in any floor expression, either
2120 because such an expression was present in the original weight function
2121 or because it was introduced when another variable with an affine bound
2122 in $x$ was summed, then the domain has to
2123 be ``splintered'' into $D$ parts, where $D$ is the least common
2124 multiple of the denominators of the coefficients of $x$ in
2125 any of the integer parts.
2126 In particular, the domain is split into $x = D y + i$ for
2127 each $i$ in $[0, D-1]$. Since $D$ is proportional to
2128 the coefficients in the constraints, it is exponential in
2129 the input size. This splintering will therefore
2130 introduce exponential behavior, even if the dimension is fixed.
2132 Splintering is clearly the most expensive step in the algorithm,
2133 so we want to avoid this step as much as possible.
2134 \shortciteN{Pugh94counting} already noted that summation should
2135 proceed over variables with integer bounds first.
2136 This can be extended to choosing a variable with the smallest
2137 coefficient in absolute value. In this way, we can avoid
2138 splintering on the largest denominator.
2140 \shortciteN{Sakellariou1996phd} claims that splintering
2141 can be avoided altogether.
2142 In particular, \shortciteN[Lemma~3.2]{Sakellariou1996phd}
2143 shows that
2145 \sum_{x=0}^a x^m\left(x \bmod b\right)^n
2148 with $a$ and $b$ integers, is equal to
2149 \begin{equation}
2150 \label{eq:3.2}
2151 \begin{cases}
2152 \displaystyle
2153 \sum_{x=0}^a x^{m+n} & \text{if $a<b$}
2155 \displaystyle
2156 \sum_{i=0}^{\floor{a/b}-1} \sum_{x=0}^{b-1} (x+ib)^m x^n +
2157 \sum_{x=0}^{a \bmod b} (x+b\floor{a/b})^m x^n & \text{if $a\ge b$}
2159 \end{cases}
2160 \end{equation}
2161 effectively avoiding splintering if a given monomial contains
2162 a single integer part expression with argument of the form
2163 $x/b$. An argument of the form $(x-c(\vec{\hat x}))/b$ can
2164 be handled through a variable substitution.
2165 If the argument is of the form $c x/b$, with $c \ne 1$,
2166 then \shortciteN[(3.27)]{Sakellariou1996phd} proposes to
2167 rewrite the monomial as
2168 \begin{align*}
2169 \sum_{x=0}^a (c x \bmod b)^n
2171 \sum_{x=0}^a \sum_{y=cx}^{cx} (y \bmod b)^n
2174 \sum_{x=0}^a
2175 \left(\sum_{y=0}^{cx} (y \bmod b)^n - \sum_{y=0}^{cx-1} (y \bmod b)^n\right)
2176 \end{align*}
2177 and applying \eqref{eq:3.2}.
2178 However, such an application results in an expression containing
2180 \sum_{y=0}^{cx \bmod b} y^n
2183 which in turn leads to a polynomial of degree $n+1$ in $(c x \bmod b)$,
2184 i.e., of degree one higher than the original expression.
2185 Furthermore, if the bound on $x$ is rational then $a$ itself contains
2186 a floor, which, on application of \eqref{eq:3.2}, results in
2187 a nested floor expression, blocking the application of the same
2188 rule for the next variable.
2189 Finally, the case where a monomial contains multiple floor
2190 expressions, either occurring in the input quasi-polynomial
2191 or introduced by different variables having a rational
2192 bound with a non-zero coefficient in the same variable, is not handled.
2193 Also note that if we disallow nested floor expressions,
2194 then this rule will rarely be applicable since we try to eliminate
2195 variables with integer bounds first.
2197 \subsection{Summation using local Euler-Maclaurin formula}
2198 \label{s:euler}
2200 \sindex{local}{Euler-Maclaurin formula}
2201 In this section we provide some implementation details
2202 on using \ai{local Euler-Maclaurin formula} to compute
2203 the sum of a piecewise polynomial evaluated in all integer
2204 points of a two-dimensional parametric polytope.
2205 For the theory behind these formula and a discussion
2206 of the original implementation (for non-parametric simplices),
2207 we refer to \shortciteN{Berline2006local}.
2209 In particular, consider a parametric piecewise polynomial
2210 in $n$ parameters and $m$ variables
2211 $c : \ZZ^n \to \ZZ^m \to \QQ : \vec p \mapsto c(\vec p)$,
2212 with $c(\vec p) : \ZZ^m \to \QQ : \vec x \mapsto c(\vec p)(\vec x)$
2215 c_{\vec p}(\vec x) =
2216 \begin{cases}
2217 c_1(\vec p)(\vec x) & \text{if $\vec x \in D_1(\vec p)$}
2219 \vdots
2221 c_r(\vec p)(\vec x) & \text{if $\vec x \in D_r(\vec p)$}
2223 \end{cases}
2225 with the $c_i$ polynomials, $c_i \in (\QQ[\vec p])[\vec x]$, and
2226 the $D_i$ disjoint linearly parametric polytopes.
2227 We want to compute
2229 g(\vec p) = \sum_{\vec x \in \ZZ^m} c(\vec p)(\vec x)
2233 \subsubsection{Reduction to the summation of a parametric polynomial
2234 over a parametric polytope with a fixed combinatorial structure}
2236 Since the $D_i$ are disjoint, we can consider each
2237 $(c_i, D_i)$-pair individually and compute
2239 g(\vec p) = \sum_{i=1}^r g_i(\vec p) =
2240 \sum_{i=1}^r \sum_{\vec x \in D_r(\vec p) \cap \ZZ^m} c_r(\vec p)(\vec x)
2243 The second step is to compute the \ai{chamber decomposition}
2244 ~\shortcite[Section 4.2.3]{Verdoolaege2005PhD} of each parametric
2245 polytope $D_i$.
2246 The result is a subdivision of the parameter space into chambers
2247 $C_{ij}$ such that $D_i$ has a fixed combinatorial structure,
2248 in particular a fixed set of parametric vertices,
2249 on (the interior of) each $C_{ij}$. Applying \autoref{p:inclusion-exclusion},
2250 this subdivision can be transformed into a partition
2251 $\{\, \tilde C_{ij} \,\}$ by
2252 making some of the facets of the chambers open%
2253 ~\shortcite[Section~3.2]{Koeppe2008parametric}.
2254 Since we are only interested in integer parameter values,
2255 any of the resulting open facets $\sp a p + c > 0$,
2256 with $\vec a \in \ZZ^n$ and $c \in \ZZ$,
2257 can then be replaced by $\sp a p + c-1 \ge 0$.
2258 Again, we have
2260 g_i(\vec p) = \sum_j g_{ij}(\vec p) =
2261 \sum_j \sum_{\vec x \in C_{ij}(\vec p) \cap \ZZ^m} c_r(\vec p)(\vec x)
2265 After this reduction, the technique of
2266 \shortciteN{Berline2006local} can be applied practically verbatim
2267 to the parametric polytope with a fixed combinatorial structure.
2268 In principle, we could also handle piecewise quasi-polynomials
2269 using the technique of \shortciteN[Section~4.5.4]{Verdoolaege2005PhD},
2270 except that we only need to create an extra variable for each
2271 distinct floor expression in a monomial, rather than for each
2272 occurrence of a floor expression in a monomial.
2273 However, since we currently only support two-dimensional polytopes,
2274 this reduction has not been implemented yet.
2276 \subsubsection{Summation over a one-dimensional parametric polytope}
2278 The basis for the summation technique is the local
2279 Euler-Maclaurin formula~\cite[Theorem~26]{Berline2006local}
2280 \begin{equation}
2281 \label{eq:EML}
2282 \sum_{\vec x \in P(\vec p) \cap \Lambda} h(\vec p)(\vec x)
2283 = \sum_{F(\vec p) \in {\mathcal F}(P(\vec p))}
2284 \int_{F(\vec p)} D_{P(\vec p),F(\vec p)} \cdot h(\vec p)
2286 \end{equation}
2287 where $P(\vec p)$ is a parametric polytope,
2288 $\Lambda$ is a lattice, ${\mathcal F}(P(\vec p))$
2289 are the faces of $P(\vec p)$, $D_{P(\vec p),F(\vec p)}$ is a
2290 specific differential operator associated to the face of a polytope.
2291 The \ai{Lebesgue measure} used in the integral is such that the
2292 integral of the indicator function of a lattice element of
2293 the lattice $\Lambda \cap (\affhull(F(\vec p)) - F(\vec p))$ is 1,
2294 i.e., the intersection of $\Lambda$ with the linear subspace
2295 parallel to the affine hull of the face $F(\vec p)$.
2296 Note that the original theorem is formulated for a non-parametric
2297 polytope and a non-parametric polynomial. However, as we will see,
2298 in each of the steps in the computation, the parameters can be
2299 treated as symbolic constants without affecting the validity of the formula,
2300 see also~\shortciteN[Section 6]{Berline2006local}.
2302 The differential operator $D_{P(\vec p),F(\vec p)}$ is obtained
2303 by plugging in the vector $\vec D=(D_1,\ldots,D_m)$ of first
2304 order differential operators, i.e., $D_k$ is the first order
2305 differential operator in the $k$th variable,
2306 in the function $\mu_{P(\vec p),F(\vec p)}$.
2307 This function is determined by the \defindex{transverse cone}
2308 of the polyhedron $P(\vec p)$ along its face $F(\vec p)$,
2309 which is the \ai{supporting cone} of $P(\vec p)$ along $F(\vec p)$
2310 projected into the linear subspace orthogonal to $F(\vec p)$.
2311 The lattice associated to this space is the projection of
2312 $\Lambda$ into this space.
2314 In particular, for a zero-dimensional affine cone in the zero-dimensional
2315 space, we have $\mu = 1$~\cite[Proposition 12]{Berline2006local},
2316 while for a one-dimensional affine
2317 cone $K = (-t + \RR_+) r$ in the one-dimensional space, where
2318 $r$ is a primitive integer vector and $t \in [0,1)$,
2319 we have~\cite[(13)]{Berline2006local}
2320 \begin{equation}
2321 \label{eq:mu:1}
2322 \mu(K)(\xi) = \frac{e^{t y}}{1-e^y} + \frac 1{y}
2323 = -\sum_{n=0}^\infty \frac{b(n+1, t)}{(n+1)!} y^n
2325 \end{equation}
2326 with $y = \sps \xi r$ and $b(n,t)$ the \ai{Bernoulli polynomial}s
2327 defined by the generating series
2329 \frac{e^{ty} y}{e^y - 1} = \sum_{n=0}^\infty \frac{b(n,t)}{n!} y^n
2332 The constant terms of these Bernoulli polynomials
2333 are the \ai{Bernoulli number}s.
2335 Applying \eqref{eq:EML} to a one-dimensional parametric polytope
2336 $P(\vec p) = [v_1(\vec p), v_2(\vec p)]$, we find
2338 \begin{aligned}
2339 \sum_{x \in P(\vec p) \cap \ZZ} h(\vec p)(x)
2340 = & \int_{P(\vec p)} D_{P(\vec p), P(\vec p)} \cdot h(\vec p)
2342 & + \int_{v_1(\vec p)} D_{P(\vec p), v_1(\vec p)} \cdot h(\vec p)
2344 & + \int_{v_2(\vec p)} D_{P(\vec p), v_2(\vec p)} \cdot h(\vec p)
2346 \end{aligned}
2348 The transverse cone of a polytope along the whole polytope is
2349 a zero-dimensional cone in a zero-dimensional space and so
2350 $D_{P(\vec p), P(\vec p)} = \mu_{P(\vec p), P(\vec p)}(D) = 1$.
2351 The transverse cone along $v_1(\vec p)$ is $v_1(\vec p) + \RR_+$
2352 and so $D_{P(\vec p), v_1(\vec p)} = \mu(v_1(\vec p) + \RR_+)(D)$
2353 as in \eqref{eq:mu:1}, with $y = \sps D 1 = D$ and
2354 $t = \ceil{v_1(\vec p)} - v_1(\vec p) =
2355 \fractional{-v_1(\vec p)}$.
2356 Similarly we find
2357 $D_{P(\vec p), v_2(\vec p)} = \mu(v_2(\vec p) - \RR_+)(D)$
2358 as in \eqref{eq:mu:1}, with $y = \sps D {-1} = -D$ and
2359 $t = v_2(\vec p) - \floor{v_2(\vec p)} =
2360 \fractional{v_2(\vec p)}$.
2361 Summarizing, we find
2363 \begin{aligned}
2364 \sum_{x \in P(\vec p) \cap \ZZ} h(\vec p)(x)
2365 = & \int_{v_1(\vec p)}^{v_2(\vec p)} h(\vec p)(t) \, dt
2367 & -\sum_{n=0}^\infty \frac{b(n+1, \fractional{-v_1(\vec p)})}{(n+1)!}
2368 (D^n h(\vec p))(v_1(\vec p))
2370 & -\sum_{n=0}^\infty (-1)^n \frac{b(n+1, \fractional{v_2(\vec p)})}{(n+1)!}
2371 (D^n h(\vec p))(v_2(\vec p))
2373 \end{aligned}
2376 Note that in order to apply this formula, we need to verify
2377 first that $v_1(\vec p)$ is indeed smaller than (or equal to)
2378 $v_2(\vec p)$. Since the combinatorial structure of $P(\vec p)$
2379 does not change throughout the interior of the chamber, we only
2380 need to check the order of the two vertices for one value
2381 of the parameters from the interior of the chamber, a point
2382 which we may compute as in \autoref{s:interior}.
2384 \subsubsection{Summation over a two-dimensional parametric polytope}
2386 For two-dimensional polytope, formula~\eqref{eq:EML} has three kinds
2387 of contributions: the integral of the polynomial over the polytope,
2388 contributions along edges and contributions along vertices.
2389 As suggested by~\citeN{Berline2007personal}, the integral can be computed
2390 by applying the Green-Stokes theorem:
2392 \iint_{P(\vec p)}
2393 \left(\frac{\partial M}{\partial x} - \frac{\partial L}{\partial y}\right) =
2394 \int_{\partial P(\vec p)} (L\, dx + M\, dy)
2397 In particular, if $M(\vec p)(x,y)$ is such that
2398 $\frac{\partial M}{\partial x}(\vec p)(x,y) = h(\vec p)(x,y)$
2399 then
2401 \iint_{P(\vec p)} h(\vec p)(x,y) =
2402 \int_{\partial P(\vec p)} M(\vec p)(x,y) \, dy
2405 Care must be taken to integrate over the boundary in the positive
2406 direction. Assuming the vertices of the polygon are not given
2407 in a predetermined order, we can check the correct orientation
2408 of the vertices of each edge individually. Let $\vec n = (n_1, n_2)$
2409 be the inner normal of a facet and let $\vec v_1(\vec p)$
2410 and $\vec v_2(\vec p)$ be the two vertices of the facet, then
2411 the vertices are in the correct order if
2413 \begin{vmatrix}
2414 v_{2,1}(\vec p)-v_{1,1}(\vec p) & n_1
2416 v_{2,2}(\vec p)-v_{1,2}(\vec p) & n_2
2417 \end{vmatrix}
2418 \ge 0
2421 Since these two vertices belong to the same edge, their order
2422 will not change within a chamber and so we can again perform
2423 this check for a single value of the parameters.
2424 To integrate $M$ over an edge $F$, let $\vec f$ be a primitive
2425 integer vector in the direction of the edge.
2426 Then $\vec v_2(\vec p) = \vec v_1(\vec p) + k(\vec p) \, \vec f$
2427 and any point on the edge can be written as
2428 $\vec v_1(\vec p) + \lambda \vec f$ with
2429 $0 \le \lambda \le k(\vec p)$.
2430 That is,
2431 \begin{equation}
2432 \label{eq:EML:int}
2433 \int_F M(\vec p)(x,y) \, dy
2435 \int_0^{k(\vec p)}
2436 M(\vec p)(v_{1,1}(\vec p) + \lambda f_1,
2437 v_{1,2}(\vec p) + \lambda f_2)
2438 f_2 \, d\lambda
2440 \end{equation}
2442 For the edges, we can again apply \eqref{eq:mu:1}, but we
2443 must first project the supporting cone at the edge into
2444 the linear subspace orthogonal to the edge.
2445 Let $\vec n = (n_1, n_2)$ be the (primitive integer) inner normal
2446 of this facet $F(\vec p)$, then $\vec f = (-n_2, n_1)$ is parallel
2447 to the facet and we can write one of the vertices $\vec v(\vec p)$
2448 as a linear combination of these two vectors:
2449 \begin{equation}
2450 \label{eq:EML:facet:coordinates}
2451 \vec v(\vec p)
2453 \begin{bmatrix}
2454 \vec f & \vec n
2455 \end{bmatrix}
2456 \vec a(\vec p)
2458 \begin{bmatrix}
2459 -n_2 & n_1 \\
2460 n_1 & n_2
2461 \end{bmatrix}
2462 \vec a(\vec p)
2463 \end{equation}
2465 \begin{equation}
2466 \label{eq:EML:facet:coordinates:2}
2467 \vec a(\vec p)
2469 \begin{bmatrix}
2470 -n_2 & n_1 \\
2471 n_1 & n_2
2472 \end{bmatrix}^{-1}
2473 \vec v(\vec p)
2475 \begin{bmatrix}
2476 -n_2/d & n_1/d \\
2477 n_1/d & n_2/d
2478 \end{bmatrix}
2479 \vec v(\vec p),
2480 \end{equation}
2481 with $d = n_1^2+n_2^2$.
2482 The lattice associated to the linear subspace orthogonal
2483 to the facet is the projection of $\Lambda$ into this space.
2484 Since $\vec n$ is primitive, a basis for this lattice can be
2485 identified with $\vec n/d$.
2486 The coordinate of the whole facet in this space is therefore
2488 d a_2(\vec p) =
2489 \begin{bmatrix}
2490 n_1 & n_2
2491 \end{bmatrix}
2492 \vec v(\vec p)
2493 $, while the transverse cone is $d a_2(\vec p) + \RR_+$.
2494 Similarly, a linear functional $\vec \xi'$ projects onto
2495 a linear functional $\xi = \sp {\xi'} n/d$ in the linear subspace.
2496 Applying \eqref{eq:mu:1}, with $y = \frac{n_1}d D_1 + \frac{n_2}d D_2$
2497 and $t = \fractional{- n_1 v_1(\vec p) - n_2 v_2(\vec p)}$, we therefore
2498 find
2499 \begin{align*}
2500 D_{P(\vec p), F(\vec p)}
2502 -\sum_{n=0}^\infty
2503 \frac{b(n+1, \fractional{-n_1 v_1(\vec p) - n_2 v_2(\vec p)})}{(n+1)!}
2504 \left(\frac{n_1}d D_1 + \frac{n_2}d D_2\right)^n
2507 - \sum_{i=0}^\infty \sum_{j=0}^\infty
2508 \frac{b(i+j+1, \fractional{-n_1 v_1(\vec p) - n_2 v_2(\vec p)})}{(i+j+1)!}
2509 \frac{n_1^i n_2^j}{d^{i+j}} D_1^i D_2^j
2511 \end{align*}
2512 After applying this differential operator to the polynomial
2513 $h(\vec p)(\vec x)$, the resulting polynomial
2515 h'(\vec p)(\vec x) = D_{P(\vec p), F(\vec p)} \cdot h(\vec p)(\vec x)
2517 needs to be integrated over the facet.
2518 The measure to be used is such that the integral of a lattice tile
2519 in the linear space parallel to the facet is 1, i.e.,
2521 \int_{\vec 0}^{\vec f} 1 = \int_0^1 1 dz = 1,
2523 with $z$ the coordinate along $\vec f$.
2524 Referring to \eqref{eq:EML:facet:coordinates} and
2525 \eqref{eq:EML:facet:coordinates:2}, all points of the facet
2526 have the form $\vec x(\vec p) = z \, \vec f + a_2(\vec p) \, \vec n$,
2527 while the $z$-coordinate of the vertices $\vec v_1(\vec p)$
2528 and $\vec v_2(\vec p)$ are
2529 $(-n_2 v_{1,1} + n_1 v_{1,2})/d$
2531 $(-n_2 v_{2,1} + n_1 v_{2,2})/d$, respectively.
2532 That is, the contribution of the facet is equal to
2534 \int_{(-n_2 v_{1,1} + n_1 v_{1,2})/d}^{(-n_2 v_{2,1} + n_1 v_{2,2})/d}
2535 h'(\vec p)\left(z \, \vec f + a_2(\vec p) \, \vec n\right) \, dz
2538 where, again, we need to ensure that the lower limit is smaller
2539 than the upper limit using the usual method of plugging in a
2540 particular value of the parameters.
2542 Finally, we consider the contributions of the vertices.
2543 The \ai{transverse cone}s are in this case simply the supporting cones.
2544 Since $\mu$ is a valuation, we may apply \ai{Barvinok's decomposition}
2545 and assume that the cone is unimodular.
2546 For an affine cone
2547 \begin{align*}
2548 K &= \vec v(\vec p) + \RR_+ \vec r_1 + \RR_+ \vec r_2
2550 &= (a_1(\vec p) + \RR_+) \vec r_1 + (a_2(\vec p) + \RR_+) \vec r_2,
2551 \end{align*}
2552 with
2554 \vec a(\vec p) =
2555 \begin{bmatrix}
2556 \vec r_1 & \vec r_2
2557 \end{bmatrix}^{-1}
2558 \vec v(\vec p)
2561 we have~\cite[Proposition~31]{Berline2006local},
2562 \begin{equation}
2563 \label{eq:mu:2}
2564 \mu(K)(\vec\xi)
2566 \frac{e^{t_1 y_1 + t_2 y_2}}{(1-e^{y_1})(1-e^{y_2})}
2567 + \frac 1{y_1}B(y_2 - C_1 y_1, t_2)
2568 + \frac 1{y_2}B(y_1 - C_2 y_2, t_1)
2569 - \frac 1{y_1 y_2},
2570 \end{equation}
2571 with
2573 B(y,t) =
2574 \frac{e^{t y}}{1-e^y} + \frac 1{y}
2575 = -\sum_{n=0}^\infty \frac{b(n+1, t)}{(n+1)!} y^n
2578 $y_i = \sps{\vec\xi}{\vec r_i}$,
2579 $C_i = \sps{\vec v_1}{\vec v_2}/\sps{\vec v_i}{\vec v_i}$
2581 $t_i = \fractional{-a_i(\vec p)}$.
2582 Expanding \eqref{eq:mu:2}, we find
2583 \begin{align*}
2584 \mu(K)(\vec\xi)
2586 \left(
2587 -\frac{b(0,t1)}{y_1} - \sum_{n=0}^\infty \frac{b(n+1,t_1)}{(n+1)!} y_1^n
2588 \right)
2589 \left(
2590 -\frac{b(0,t2)}{y_2} - \sum_{n=0}^\infty \frac{b(n+1,t_2)}{(n+1)!} y_2^n
2591 \right)
2593 & \phantom{=}
2595 \left(
2596 \sum_{n=0}^\infty \frac{b(n+1,t_2)}{(n+1)!} \frac{y_2^n}{y_1}
2598 \sum_{n=0}^\infty \frac{b(n+1,t_2)}{(n+1)!} \frac{(y_2-C_1 y_1)^n-y_2^n}{y_1}
2599 \right)
2601 & \phantom{=}
2603 \left(
2604 \sum_{n=0}^\infty \frac{b(n+1,t_1)}{(n+1)!} \frac{y_1^n}{y_2}
2606 \sum_{n=0}^\infty \frac{b(n+1,t_1)}{(n+1)!} \frac{(y_1-C_2 y_2)^n-y_1^n}{y_2}
2607 \right)
2609 & \phantom{=}
2610 - \frac 1{y_1 y_2}
2613 \sum_{n_1=0}^\infty
2614 \sum_{n_2=0}^\infty
2615 c(C_1, C_2, t_1, t_2; n_1, n_2) \, y_1^n y_2^n
2617 \end{align*}
2618 with
2619 \begin{align*}
2620 c(C_1, C_2, t_1, t_2; n_1, n_2)
2622 \frac{b(n_1+1,t_1)}{(n_1+1)!} \frac{b(n_2+1,t_2)}{(n_2+1)!}
2626 \frac{b(n_1+n_2+1,t_2)}{(n_1+n_2+1)!} {n_1+n_2+1 \choose n_1+1}
2627 \left(-C_1\right)^{n_1+1}
2631 \frac{b(n_1+n_2+1,t_1)}{(n_1+n_2+1)!} {n_1+n_2+1 \choose n_2+1}
2632 \left(-C_2\right)^{n_2+1}
2634 \end{align*}
2635 For $\vec \xi = \vec D$, we have
2636 \begin{align*}
2637 y_1^n y_2^n
2639 \left( r_{1,1} D_1 + r_{1,2} D_2 \right)^{n_1}
2640 \left( r_{2,1} D_1 + r_{2,2} D_2 \right)^{n_2}
2643 \left(
2644 \sum_{k=0}^{n_1} r_{1,1}^k r_{1,2}^{n_1 - k} { n_1 \choose k} D_1^k D_2^{n_1-k}
2645 \right)
2646 \left(
2647 \sum_{l=0}^{n_2} r_{2,1}^l r_{2,2}^{n_2 - l} { n_2 \choose l} D_1^l D_2^{n_2-l}
2648 \right)
2649 \end{align*}
2650 and so
2652 D_{P(\vec p), \vec v(\vec p)} = \mu(K)(\vec D)
2656 \sum_{i=0}^\infty
2657 \sum_{j=0}^\infty
2658 \sum_{\shortstack{$\scriptstyle i+j = n_1+n_2$\\$\scriptstyle n_1 \ge 0$\\$\scriptstyle n_2 \ge 0$}}
2659 \sum_{\shortstack{$\scriptstyle k+l = i$\\$\scriptstyle 0 \le k \le n_1$\\$\scriptstyle 0 \le l \le n_2$}}
2660 c(C_1, C_2, t_1, t_2; n_1, n_2)
2661 r_{1,1}^k r_{1,2}^{n_1 - k}
2662 r_{2,1}^l r_{2,2}^{n_2 - l}
2663 { n_1 \choose k} { n_2 \choose l} D_1^i D_2^j
2666 The contribution of this vertex is then
2668 h'(\vec p)(\vec v(\vec p))
2671 with $
2672 h'(\vec p)(\vec x) = D_{P(\vec p), \vec v(\vec p)} \cdot h(\vec p)(\vec x)
2675 \begin{example}
2676 As a simple example, consider the (non-parametric) triangle
2677 in \autoref{f:EML:triangle} and assume we want to compute
2679 \sum_{\vec x \in T \cap \ZZ^2} x_1 x_2
2682 Since $T \cap \ZZ^2 = \{\, (2,4), (3,4), (2,5) \, \}$,
2683 the result should be
2685 2 \cdot 4 + 3 \cdot 4 + 2 \cdot 5 = 30
2689 \begin{figure}
2690 \intercol=1.2cm
2691 \begin{xy}
2692 <\intercol,0pt>:<0pt,\intercol>::
2693 \POS@i@={(2,4),(3,4),(2,5),(2,4)},{0*[|(2)]\xypolyline{}}
2694 \POS(2.35,4.25)*{x_1 x_2}
2695 \POS(2,4)*+!U{(2,4)}
2696 \POS(3,4)*+!U{(3,4)}
2697 \POS(2,5)*+!D{(2,5)}
2698 \POS(2,4)*{\cdot}
2699 \POS(3,4)*{\cdot}
2700 \POS(2,5)*{\cdot}
2701 \POS(-1,0)\ar(4,0)
2702 \POS(0,-1)\ar(0,5.5)
2703 \end{xy}
2704 \caption{Sum of polynomial $x_1 x_2$ over the integer points in a triangle $T$}
2705 \label{f:EML:triangle}
2706 \end{figure}
2708 Let us first consider the integral
2710 \iint_T x_1 x_2 = \int_{\partial T} \frac{x_1^2 x_2}2 \, d x_2
2713 Integration along each of the edges of the triangle yields
2714 the following.
2716 \marginpar{%
2717 \intercol=1cm
2718 \begin{xy}
2719 <\intercol,0pt>:<0pt,\intercol>::
2720 \POS(0,-1)*\xybox{
2721 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2722 \POS(0,0)\ar@[|(2)](1,0)
2724 \end{xy}
2726 For the edge in the margin, we have $\vec f = (1,0)$, i.e., $f_2 = 0$.
2727 The contribution of this edge to the integral is therefore zero.
2729 \marginpar{%
2730 \intercol=1cm
2731 \begin{xy}
2732 <\intercol,0pt>:<0pt,\intercol>::
2733 \POS(0,-1)*\xybox{
2734 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2735 \POS(1,0)\ar@[|(2)](0,1)
2737 \end{xy}
2739 For this edge, we have $\vec f = (-1,1)$.
2740 The contribution of this edge to the integral is therefore
2742 \int_0^1 \frac{(3-\lambda)^2(4+\lambda)}2 d\lambda
2743 = \frac{337}{24}
2747 \marginpar{%
2748 \intercol=1cm
2749 \begin{xy}
2750 <\intercol,0pt>:<0pt,\intercol>::
2751 \POS(0,-1)*\xybox{
2752 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2753 \POS(0,1)\ar@[|(2)](0,0)
2755 \end{xy}
2757 For this edge, we have $\vec f = (0,-1)$.
2758 The contribution of this edge to the integral is therefore
2760 \int_0^1 \frac{2^2(5-\lambda)}2 (-1) d\lambda
2761 = -9
2765 The total integral is therefore
2767 \int_{\partial T} \frac{x_1^2 x_2}2 \, d x_2
2768 = 0 + \frac{337}{24} - 9 = \frac{121}{24}
2772 Now let us consider the contributions of the edges.
2773 We will need the following \ai{Bernoulli number}s in our
2774 computations.
2775 \begin{align*}
2776 b(1,0) & = - \frac 1 2
2778 b(2,0) & = \frac 1 6
2780 b(3,0) & = 0
2782 b(4,0) & = -\frac 1 {30}
2783 \end{align*}
2785 \marginpar{%
2786 \intercol=1cm
2787 \begin{xy}
2788 <\intercol,0pt>:<0pt,\intercol>::
2789 \POS(0,-1)*\xybox{
2790 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2791 \POS(0,0)\ar@[|(2)]@{-}(1,0)
2792 \POS(0.5,0)\ar(0.5,1)
2794 \end{xy}
2796 The normal to the facet $F_1$ in the margin is $\vec n = (0,1)$.
2797 The vector $\vec f = (-1,0)$ is parallel to the facet.
2798 We have
2800 \begin{bmatrix}
2801 2 \\ 4
2802 \end{bmatrix}
2805 \begin{bmatrix}
2806 -1 \\ 0
2807 \end{bmatrix}
2809 \begin{bmatrix}
2810 0 \\ 1
2811 \end{bmatrix}
2812 \quad\text{and}\quad
2813 \begin{bmatrix}
2814 3 \\ 4
2815 \end{bmatrix}
2818 \begin{bmatrix}
2819 -1 \\ 0
2820 \end{bmatrix}
2822 \begin{bmatrix}
2823 0 \\ 1
2824 \end{bmatrix}
2827 Therefore $t = \fractional{-4} = 0$, $y = D_2$,
2828 \begin{align*}
2829 D_{T,F_1}
2830 & =
2831 - \sum_{j=0}^\infty \frac{b(j+1, 0)}{(j+1)!} D_2^j
2834 - \frac{b(1,0)}1 - \frac{b(2,0)}2 D_2 + \cdots
2835 \end{align*}
2838 h'(\vec x) =
2839 D_{T,F_1} \cdot x_1 x_2 =
2840 \left(\frac 1 2 - \frac 1{12} D_2\right) \cdot x_1 x_2
2842 \frac 1 2 x_1 x_2 - \frac 1{12} x_1
2845 With $x_1 = - z$ and $x_2 = 4$, the contribution of this facet
2848 \int_{-3}^{-2} - 2 z + \frac 1{12} z \, dz
2850 \frac{115}{24}
2854 \marginpar{%
2855 \intercol=1cm
2856 \begin{xy}
2857 <\intercol,0pt>:<0pt,\intercol>::
2858 \POS(0,-1)*\xybox{
2859 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2860 \POS(0,0)\ar@[|(2)]@{-}(0,1)
2861 \POS(0,0.5)\ar(1,0.5)
2863 \end{xy}
2865 The normal to the facet $F_2$ in the margin is $\vec n = (1,0)$.
2866 The vector $\vec f = (0,1)$ is parallel to the facet.
2867 We have
2869 \begin{bmatrix}
2870 2 \\ 4
2871 \end{bmatrix}
2874 \begin{bmatrix}
2875 0 \\ 1
2876 \end{bmatrix}
2878 \begin{bmatrix}
2879 1 \\ 0
2880 \end{bmatrix}
2881 \quad\text{and}\quad
2882 \begin{bmatrix}
2883 2 \\ 5
2884 \end{bmatrix}
2887 \begin{bmatrix}
2888 0 \\ 1
2889 \end{bmatrix}
2891 \begin{bmatrix}
2892 1 \\ 0
2893 \end{bmatrix}
2896 Therefore $t = \fractional{-2} = 0$, $y = D_1$,
2897 \begin{align*}
2898 D_{T,F_2}
2899 & =
2900 - \sum_{i=0}^\infty \frac{b(i+1, 0)}{(i+1)!} D_1^i
2903 - \frac{b(1,0)}1 - \frac{b(2,0)}2 D_1 + \cdots
2904 \end{align*}
2907 h'(\vec x) =
2908 D_{T,F_2} \cdot x_1 x_2 =
2909 \left(\frac 1 2 - \frac 1{12} D_1\right) \cdot x_1 x_2
2911 \frac 1 2 x_1 x_2 - \frac 1{12} x_2
2914 With $x_1 = 2$ and $x_2 = z$, the contribution of this facet
2917 \int_{4}^{5} z - \frac 1{12} z \, dz
2919 \frac{33}{8}
2923 \marginpar{%
2924 \intercol=1cm
2925 \begin{xy}
2926 <\intercol,0pt>:<0pt,\intercol>::
2927 \POS(0,-1)*\xybox{
2928 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2929 \POS(1,0)\ar@[|(2)]@{-}(0,1)
2930 \POS(0.5,0.5)\ar(-0.5,-0.5)
2932 \end{xy}
2934 The normal to the facet $F_3$ in the margin is $\vec n = (-1,-1)$.
2935 The vector $\vec f = (1,-1)$ is parallel to the facet.
2936 We have
2938 \begin{bmatrix}
2939 3 \\ 4
2940 \end{bmatrix}
2942 -\frac 1 2
2943 \begin{bmatrix}
2944 1 \\ -1
2945 \end{bmatrix}
2946 -\frac 7 2
2947 \begin{bmatrix}
2948 -1 \\ -1
2949 \end{bmatrix}
2950 \quad\text{and}\quad
2951 \begin{bmatrix}
2952 2 \\ 5
2953 \end{bmatrix}
2955 -\frac 3 2
2956 \begin{bmatrix}
2957 1 \\ -1
2958 \end{bmatrix}
2959 -\frac 7 2
2960 \begin{bmatrix}
2961 -1 \\ -1
2962 \end{bmatrix}
2965 Therefore $t = \fractional{7} = 0$, $y = -\frac 1 2 D_1 -\frac 1 2 D_2$,
2966 \begin{align*}
2967 D_{T,F_3}
2968 & =
2969 - \sum_{i=0}^\infty \sum_{j=0}^\infty
2970 \frac{b(i+j+1, 0)}{(i+j+1)!}
2971 \frac{(-1)^{i+j}}{2^{i+j}} D_1^i D_2^j
2974 - \frac{b(1,0)}1
2975 + \frac 1 2 \frac{b(2,0)}2 D_1
2976 + \frac 1 2 \frac{b(2,0)}2 D_2 + \cdots
2977 \end{align*}
2980 h'(\vec x) =
2981 D_{T,F_4} \cdot x_1 x_2 =
2982 \left(\frac 1 2 + \frac 1{24} D_1 + \frac 1{24} D_2\right) \cdot x_1 x_2
2984 \frac 1 2 x_1 x_2 + \frac 1{24} x_2 + \frac 1{24} x_1
2987 With $x_1 = z + \frac 7 2$ and $x_2 = -z + \frac 7 2$,
2988 the contribution of this facet
2991 \int_{-\frac 3 2}^{-\frac 1 2}
2992 \frac 1 2 (z + \frac 7 2)(-z + \frac 7 2)
2993 + \frac 1{24}(-z + \frac 7 2)
2994 + \frac 1{24}(z + \frac 7 2) \, dz
2996 \frac{47}{8}
3000 The total contribution of the edges is therefore
3002 \frac{115}{24}+\frac{33}8+
3003 \frac{47}{8} = \frac{355}{24}
3007 Finally, we consider the contributions of the vertices.
3009 \marginpar{%
3010 \intercol=1cm
3011 \begin{xy}
3012 <\intercol,0pt>:<0pt,\intercol>::
3013 \POS(0,-1)*\xybox{
3014 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
3015 \POS(1,0)\ar@[|(2)](0,1)
3016 \POS(1,0)\ar@[|(2)](0,0)
3018 \end{xy}
3020 For the vertex $\vec v = (3,4)$, we have
3021 $\vec r_1 = (-1,0)$ and $\vec r_2 = (-1,1)$.
3022 Since $\vec v$ is integer, we have $t_1 = t_2 = 0$.
3023 Also, $C_1 = 1$, $C_2 = 1/2$, $y_1 = -D_1$ and $y_2 = -D_1 + D_2$.
3024 Since the total degree of the polynomial $x_1 x_2$ is two,
3025 we only need the coefficients of $\mu(K)(\vec \xi)$ up to
3026 $n_1+n_2 = 2$.
3028 \noindent
3029 \begin{tabular}{c|c|c}
3030 $n_1$ & $n_2$
3032 \hline
3033 0 & 0 &
3035 \left(
3036 \frac{b(1,0)}{1!}
3037 \frac{b(1,0)}{1!}
3039 \frac{b(2,0)}{2!}
3040 {1 \choose 1}(-1)^1
3042 \frac{b(2,0)}{2!}
3043 {1 \choose 1}(-\frac 12)^1
3044 \right)
3047 1 & 0 &
3049 \left(
3050 \frac{b(2,0)}{2!}
3051 \frac{b(1,0)}{1!}
3053 \frac{b(3,0)}{3!}
3054 {2 \choose 2}(-1)^2
3056 \frac{b(3,0)}{3!}
3057 {2 \choose 1}(-\frac 12)^1
3058 \right)
3059 \left(
3060 -D_1
3061 \right)
3064 0 & 1 &
3066 \left(
3067 \frac{b(1,0)}{1!}
3068 \frac{b(2,0)}{2!}
3070 \frac{b(3,0)}{3!}
3071 {2 \choose 1}(-1)^1
3073 \frac{b(3,0)}{3!}
3074 {2 \choose 2}(-\frac 12)^2
3075 \right)
3076 \left(
3077 -D_1 + D_2
3078 \right)
3081 2 & 0 &
3083 \left(
3084 \frac{b(3,0)}{3!}
3085 \frac{b(1,0)}{1!}
3087 \frac{b(4,0)}{4!}
3088 {3 \choose 3}(-1)^3
3090 \frac{b(4,0)}{4!}
3091 {3 \choose 1}(-\frac 12)^1
3092 \right)
3093 \left(
3094 -D_1
3095 \right)^2
3098 1 & 1 &
3100 \left(
3101 \frac{b(2,0)}{2!}
3102 \frac{b(2,0)}{2!}
3104 \frac{b(4,0)}{4!}
3105 {3 \choose 2}(-1)^2
3107 \frac{b(4,0)}{4!}
3108 {3 \choose 2}(-\frac 12)^2
3109 \right)
3110 \left(
3111 -D_1
3112 \right)
3113 \left(
3114 -D_1 + D_2
3115 \right)
3118 0 & 2 &
3120 \left(
3121 \frac{b(1,0)}{1!}
3122 \frac{b(3,0)}{3!}
3124 \frac{b(4,0)}{4!}
3125 {3 \choose 1}(-1)^1
3127 \frac{b(4,0)}{4!}
3128 {3 \choose 3}(-\frac 12)^3
3129 \right)
3130 \left(
3131 -D_1 + D_2
3132 \right)^2
3134 \end{tabular}
3136 We find
3137 \begin{align*}
3138 h'(\vec x)
3140 \left(
3141 \frac 3 8 - \frac 1{24} (-D_1) - \frac 1{24} (-D_1 + D_2)
3142 + \frac 7{576} (-D_1 D_2)
3143 - \frac 5{1152} (-2 D_1 D2)
3144 \right) x_1 x_2
3147 \frac 3 8 x_1 x_2 + \frac 1{24} x_2 - \frac 1{24} (-x_2 + x_1)
3148 + \frac 7{576} (-1)
3149 - \frac 5{1152} (-2)
3151 \end{align*}
3152 The contribution of this vertex is therefore
3154 h'(3,4) = \frac {1355}{288}
3158 \marginpar{%
3159 \intercol=1cm
3160 \begin{xy}
3161 <\intercol,0pt>:<0pt,\intercol>::
3162 \POS(0,-1)*\xybox{
3163 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
3164 \POS(0,1)\ar@[|(2)](1,0)
3165 \POS(0,1)\ar@[|(2)](0,0)
3167 \end{xy}
3169 For the vertex $\vec v = (2,5)$, we have
3170 $\vec r_1 = (0,-1)$ and $\vec r_2 = (1,-1)$.
3171 Since $\vec v$ is integer, we have $t_1 = t_2 = 0$.
3172 Also, $C_1 = 1$, $C_2 = 1/2$, $y_1 = -D_2$ and $y_2 = D_1 - D_2$.
3173 We similarly find
3175 h'(\vec x)
3177 \frac 3 8 x_1 x_2 + \frac 1{24} x_1 - \frac 1{24} (x_2 - x_1)
3178 + \frac 7{576} (-1)
3179 - \frac 5{1152} (-2)
3182 The contribution of this vertex is therefore
3184 h'(2,5) = \frac {1067}{288}
3188 \marginpar{%
3189 \intercol=1cm
3190 \begin{xy}
3191 <\intercol,0pt>:<0pt,\intercol>::
3192 \POS(0,-1)*\xybox{
3193 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
3194 \POS(0,0)\ar@[|(2)](1,0)
3195 \POS(0,0)\ar@[|(2)](0,1)
3197 \end{xy}
3199 For the vertex $\vec v = (2,4)$, we have
3200 $\vec r_1 = (1,0)$ and $\vec r_2 = (0,1)$.
3201 Since $\vec v$ is integer, we have $t_1 = t_2 = 0$.
3202 The computations are easier in this case since
3203 $C_1 = C_2 = 0$, $y_1 = D_1$ and $y_2 = D_2$.
3204 We find
3206 h'(\vec x)
3208 \frac 1 4 x_1 x_2 - \frac 1{12} x_2 - \frac 1{12} x_1
3209 + \frac 1{144} (1)
3212 The contribution of this vertex is therefore
3214 h'(2,4) = \frac {253}{144}
3218 The total contribution of the vertices is then
3220 \frac {1355}{288} + \frac {1067}{288} + \frac {253}{144}
3221 = \frac {61}6
3223 and the total sum is
3225 \frac{121}{24}+\frac{355}{24}+\frac{61}6 = 30
3229 \end{example}
3231 \begin{example}
3232 Consider the parametric polytope
3234 P(n) = \{\, \vec x \mid x_1 \ge 2 \wedge 3 x_1 \le n + 9
3235 \wedge 4 \le x_2 \le 5 \,\}
3238 If $n \ge -3$, then the vertices of this polytope are
3239 $(2,4)$, $(2,5)$, $(3+n/3,4)$ and $(3+n/3,5)$.
3240 The contributions of the faces of $P(n)$ to
3242 \sum_{\vec x \in P(n) \cap \ZZ^2} x_1 x_2
3244 for the chamber $n \ge -3$ are shown in \autoref{t:sum:rectangle}.
3245 The final result is
3247 \begin{cases}
3248 \frac{ n^2}{2}
3249 - 3 n \fractional{\frac{ n}{3}}
3250 + \frac{21}{2} n
3251 + \frac{9}{2} \fractional{\frac{ n}{3}}^2
3252 - \frac{63}{2} \fractional{\frac{ n}{3}}
3253 + 45
3254 & \text{if $ n+3 \ge 0$}.
3255 \end{cases}
3258 \begin{table}
3259 \intercol=1cm
3260 \begin{tabular}{lc}
3261 \begin{xy}
3262 <\intercol,0pt>:<0pt,\intercol>::
3263 \POS(-1,-0.5)*\xybox{
3264 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*[|(2)]\xypolyline{}}
3265 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3267 \end{xy}
3270 \displaystyle
3271 \frac{ n^2}{4}
3272 + \frac{9}{2} n
3273 + \frac{45}{4}
3276 \begin{xy}
3277 <\intercol,0pt>:<0pt,\intercol>::
3278 \POS(-1,-0.5)*\xybox{
3279 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3280 \POS(0,0)\ar@[|(2)]@{-}(0,1)
3281 \POS(0,0.5)*+!L{2}
3282 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3284 \end{xy}
3287 \displaystyle
3288 \frac{33}{8}
3291 \begin{xy}
3292 <\intercol,0pt>:<0pt,\intercol>::
3293 \POS(-1,-0.5)*\xybox{
3294 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3295 \POS(1,0)\ar@[|(2)]@{-}(1,1)
3296 \POS(1,0.5)*+!L{3+n/3}
3297 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3299 \end{xy}
3302 \displaystyle
3303 - \frac{3}{2} n \fractional{\frac{ n}{3}}
3304 + \frac{3}{4} n
3305 + \frac{9}{4} \fractional{\frac{ n}{3}}^2
3306 - \frac{63}{4} \fractional{\frac{ n}{3}}
3307 + \frac{57}{8}
3310 \begin{xy}
3311 <\intercol,0pt>:<0pt,\intercol>::
3312 \POS(-1,-0.5)*\xybox{
3313 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3314 \POS(0,0)\ar@[|(2)]@{-}(1,0)
3315 \POS(0.5,0)*+!D{4}
3316 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3318 \end{xy}
3321 \displaystyle
3322 \frac{23}{216} n^2
3323 + \frac{23}{12} n
3324 + \frac{115}{24}
3327 \begin{xy}
3328 <\intercol,0pt>:<0pt,\intercol>::
3329 \POS(-1,-0.5)*\xybox{
3330 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3331 \POS(0,1)\ar@[|(2)]@{-}(1,1)
3332 \POS(0.5,1)*+!U{5}
3333 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3335 \end{xy}
3338 \displaystyle
3339 \frac{31}{216} n^2
3340 + \frac{31}{12} n
3341 + \frac{155}{24}
3344 \begin{xy}
3345 <\intercol,0pt>:<0pt,\intercol>::
3346 \POS(-1,-0.5)*\xybox{
3347 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3348 \POS(1,1)\ar@[|(2)](1,0)
3349 \POS(1,1)\ar@[|(2)](0,1)
3350 \POS(1,1)*+!LU{(3+n/3,5)}
3351 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3353 \end{xy}
3356 \displaystyle
3357 - \frac{31}{36} n \fractional{\frac{ n}{3}}
3358 + \frac{31}{72} n
3359 + \frac{31}{24} \fractional{\frac{ n}{3}}^2
3360 - \frac{217}{24} \fractional{\frac{ n}{3}}
3361 + \frac{589}{144}
3364 \begin{xy}
3365 <\intercol,0pt>:<0pt,\intercol>::
3366 \POS(-1,-0.5)*\xybox{
3367 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3368 \POS(0,1)\ar@[|(2)](1,1)
3369 \POS(0,1)\ar@[|(2)](0,0)
3370 \POS(0,1)*+!LU{(2,5)}
3371 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3373 \end{xy}
3376 \displaystyle
3377 \frac{341}{144}
3380 \begin{xy}
3381 <\intercol,0pt>:<0pt,\intercol>::
3382 \POS(-1,-0.5)*\xybox{
3383 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3384 \POS(0,0)\ar@[|(2)](1,0)
3385 \POS(0,0)\ar@[|(2)](0,1)
3386 \POS(0,0)*+!LD{(2,4)}
3387 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3389 \end{xy}
3392 \displaystyle
3393 \frac{253}{144}
3396 \begin{xy}
3397 <\intercol,0pt>:<0pt,\intercol>::
3398 \POS(-1,-0.5)*\xybox{
3399 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3400 \POS(1,0)\ar@[|(2)](1,1)
3401 \POS(1,0)\ar@[|(2)](0,0)
3402 \POS(1,0)*+!LD{(3+n/3,4)}
3403 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3405 \end{xy}
3408 \displaystyle
3409 - \frac{23}{36} n \fractional{\frac{ n}{3}}
3410 + \frac{23}{72} n
3411 + \frac{23}{24} \fractional{\frac{ n}{3}}^2
3412 - \frac{161}{24} \fractional{\frac{ n}{3}}
3413 + \frac{437}{144}
3415 \end{tabular}
3416 \caption{Contributions of the faces of $P(n)$ to the sum of $x_1 x_2$ over
3417 the integer points of $P(n)$}
3418 \label{t:sum:rectangle}
3419 \end{table}
3421 \end{example}
3424 \subsection{Conversion to ``standard form''}
3425 \label{s:standard}
3427 Some algorithms or tools expect a polyhedron to be
3428 specified in ``\ai{standard form}'', i.e.,
3429 \begin{equation}
3430 \label{eq:standard}
3431 \begin{cases}
3432 \begin{aligned}
3433 A \vec x & = \vec b \\
3434 \vec x & \ge \vec 0
3436 \end{aligned}
3437 \end{cases}
3438 \end{equation}
3439 Given an arbitrary (parametric) polyhedron
3440 \begin{equation}
3441 \label{eq:non-standard}
3442 \{\,
3443 \vec x \mid
3444 A \vec x + \vec b(\vec p) \ge 0
3445 \,\}
3447 \end{equation}
3448 a conversion to standard form requires the introduction
3449 of \ai{slack variable}s and a way of dealing with variables
3450 of \ai{unrestricted sign}.
3451 In this section we will be satisfied with a reduction
3452 to the form
3453 \begin{equation}
3454 \label{eq:standard:2}
3455 \begin{cases}
3456 \begin{aligned}
3457 A \vec x & = \vec b \\
3458 D \vec x & \ge \vec c
3460 \end{aligned}
3461 \end{cases}
3462 \end{equation}
3463 with $D$ a diagonal matrix with positive entries.
3464 That is, we do not necessarily make all variables non-negative,
3465 but we do ensure that they have a lower bound.
3466 If needed, a subsequent reduction can then be performed.
3468 The standard way of dealing with variables of unrestricted
3469 sign is to replace a variable $x$ of unknown sign by the
3470 difference ($x = x' - x''$) of two non-negative variables
3471 ($x', x'' \ge 0$).
3472 However, some algorithms are somewhat sensitive with respect
3473 to the number of variables and so we would prefer to introduce
3474 as few extra variables as possible.
3475 We will therefore apply a \ai{unimodular transformation}
3476 on the variables such that all transformed variables are known
3477 to be non-negative.
3479 The first step is to compute the \indac{HNF} of A,
3480 i.e., a matrix $H = A U$, with $U$ unimodular,
3481 in column echelon form such that the
3482 first entry in each column is positive and the other entries
3483 on the corresponding row are non-negative and strictly smaller
3484 than this first entry.
3485 By reordering the rows we may assume that the top square part
3486 of $H$ is lower-triangular.
3487 By a further unimodular transformation, the entries
3488 below the diagonal can be made non-positive and strictly
3489 smaller (in absolute value) than the diagonal entry of the same row.
3491 For each of the new variables, we can take a positive
3492 combination of the corresponding row and the previous rows
3493 to obtain a positive multiple of the corresponding unit vector,
3494 implying that the variable has a lower bound.
3495 A slack variable can then be introduced for each of the
3496 rows in the top square part of $H'$ that is not already
3497 a positive multiple of a unit vector and for each of
3498 the rows below the top square part of $H'$.
3500 \begin{example}
3501 Consider the cone
3503 \left\{\,
3504 \vec x \mid
3505 \begin{bmatrix}
3506 67 & -13 \\
3507 -52 & 53
3508 \end{bmatrix}
3509 \vec x
3511 \vec 0
3512 \,\right\}
3515 This cone is already situated in the first quadrant,
3516 but this may not be obvious from the constraints.
3517 Furthermore, directly adding slack variables would
3518 lead to a total of 4 variables, whereas we can also
3519 represent this cone in standard form with only 3 variables.
3520 We have
3522 H' =
3523 \begin{bmatrix}
3524 1 & 0 \\
3525 -1331 & 2875
3526 \end{bmatrix}
3528 \begin{bmatrix}
3529 67 & -13 \\
3530 -52 & 53
3531 \end{bmatrix}
3532 \begin{bmatrix}
3533 -6 & 13 \\
3534 -31 & 57
3535 \end{bmatrix}
3536 = A U'
3539 Adding a slack variable for the second row of $H'$, we
3540 obtain the equivalent problem
3542 \begin{cases}
3543 \begin{aligned}
3544 \begin{bmatrix}
3545 -1331 & 2875 & -1
3546 \end{bmatrix}
3547 \vec x'
3549 \vec 0
3551 \vec x' & \ge \vec 0
3552 \end{aligned}
3553 \end{cases}
3555 with
3557 \vec x =
3558 \begin{bmatrix}
3559 -6 & 13 & 0 \\
3560 -31 & 57 & 0
3561 \end{bmatrix}
3562 \vec x'
3565 \end{example}
3567 A similar construction was used by \shortciteN[Lemma~3.10]{Eisenbrand2000PhD}
3568 and \shortciteN{Hung1990}.
3570 \subsection{Using TOPCOM to compute Chamber Decompositions}
3572 In this section, we describe how to use the correspondence
3573 between the \ai{regular triangulation}s of a point set
3574 and the chambers of the \ai{Gale transform}
3575 of the point set~\shortcite{Gelfand1994}
3576 to compute the chamber decomposition of a parametric polytope.
3577 This correspondence was also used by \shortciteN{Pfeifle2003}
3578 \shortciteN{Eisenschmidt2007integrally}.
3580 Let us first assume that the parametric polytope can be written as
3581 \begin{equation}
3582 \label{eq:TOPCOM:polytope}
3583 \begin{cases}
3584 \begin{aligned}
3585 \vec x &\ge 0
3587 A \, \vec x &\le \vec b(\vec p)
3589 \end{aligned}
3590 \end{cases}
3591 \end{equation}
3592 where the right hand side $\vec b(\vec p)$ is arbitrary and
3593 may depend on the parameters.
3594 The first step is to add slack variables $\vec s$ to obtain
3595 the \ai{vector partition} problem
3597 \begin{cases}
3598 \begin{aligned}
3599 A \, \vec x + I \, \vec s & = \vec b(\vec p)
3601 \vec x, \vec s &\ge 0
3603 \end{aligned}
3604 \end{cases}
3606 with $I$ the identity matrix.
3607 Then we compute the (right) kernel $K$ of the matrix
3608 $\begin{bmatrix}
3609 A & I
3610 \end{bmatrix}$, i.e.,
3612 \begin{bmatrix}
3613 A & I
3614 \end{bmatrix}
3619 and use \ai[\tt]{TOPCOM}'s \ai[\tt]{points2triangs} to
3620 compute the \ai{regular triangulation}s of the points specified
3621 by the rows of $K$.
3622 Each of the resulting triangulations corresponds to a chamber
3623 in the chamber complex of the above vector partition problem.
3624 Each simplex in a triangulation corresponds to a parametric
3625 vertex active on the corresponding chamber and
3626 each point in the simplex (i.e., a row of $K$) corresponds
3627 to a variable ($x_j$ or $s_j$) that is set to zero to obtain
3628 this parametric vertex.
3629 In the original formulation of the problem~\eqref{eq:TOPCOM:polytope}
3630 each such variable set to zero reflects the saturation of the
3631 corresponding constraint ($x_j = 0$ for $x_j = 0$ and
3632 $\sps {\vec a_j}{\vec x} = b_j(\vec p)$ for $s_j = 0$).
3633 A description of the chamber can then be obtained by plugging
3634 in the parametric vertices in the remaining constraints.
3636 \begin{example}
3637 Consider the parametric polytope
3639 P(p,q,r) = \{\,
3640 (i,j) \mid 0 \le i \le p \wedge
3641 0 \le j \le 2 i + q \wedge
3642 0 \le k \le i - p + r \wedge
3643 p \ge 0 \wedge
3644 q \ge 0 \wedge
3645 r \ge 0
3646 \,\}
3649 The constraints involving the variables are
3651 \begin{cases}
3652 \begin{aligned}
3653 \begin{bmatrix}
3658 & & 1
3659 \end{bmatrix}
3660 \begin{bmatrix}
3661 i \\ j \\ k
3662 \end{bmatrix}
3664 \begin{matrix}
3670 \end{matrix}
3671 \begin{array}{l}
3677 \end{array}
3679 \begin{bmatrix}
3680 1 & 0 & 0
3682 -1 & 0 & 1
3684 -2 & 1 & 0
3685 \end{bmatrix}
3686 \begin{bmatrix}
3687 i \\ j \\ k
3688 \end{bmatrix}
3690 \begin{matrix}
3696 \end{matrix}
3697 \begin{array}{l}
3702 -p + r
3703 \end{array}
3704 \end{aligned}
3705 \end{cases}
3707 We have
3709 \begin{bmatrix}
3710 1 & 0 & 0 & 1 & 0 & 0 \\
3711 -1 & 0 & 1 & 0 & 1 & 0 \\
3712 -2 & 1 & 0 & 0 & 0 & 1 \\
3713 \end{bmatrix}
3714 \begin{bmatrix}
3715 -1 & 0 & 0 \\
3716 -2 & 0 & -1 \\
3717 -1 & -1 & 0 \\
3718 1 & 0 & 0 \\
3719 0 & 1 & 0 \\
3720 0 & 0 & 1 \\
3721 \end{bmatrix}
3725 Computing the \ai{regular triangulation}s of the rows of $K$
3726 using \ai[\tt]{TOPCOM}, we obtain
3727 \begin{verbatim}
3728 > cat e2.topcom
3730 [ -1 0 0 ]
3731 [ -2 0 -1 ]
3732 [ -1 -1 0 ]
3733 [ 1 0 0 ]
3734 [ 0 1 0 ]
3735 [ 0 0 1 ]
3737 > points2triangs --regular < e2.topcom
3738 T[1]:={{0,1,2},{1,2,3},{0,1,4},{1,3,4},{0,2,5},{2,3,5},{0,4,5},{3,4,5}};
3739 T[2]:={{1,2,3},{1,3,4},{2,3,5},{3,4,5},{1,2,5},{1,4,5}};
3740 T[3]:={{1,2,3},{1,3,4},{2,3,5},{3,4,5},{1,2,4},{2,4,5}};
3741 \end{verbatim}
3743 We see that we have three chambers in the decomposition,
3744 one with 8 vertices and two with 6 vertices.
3745 Take the second vertex (``\verb+{1,2,3}+'') of the first chamber.
3746 This vertex corresponds
3747 to the saturation of the constraints $j \ge 0$, $k \ge 0$
3748 and $i \le p$, i.e., $(i,j,k) = (p,0,0)$. Plugging in this
3749 vertex in the remaining constraints, we see that it is only valid
3750 in case $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$.
3751 For the remaining vertices of the first chamber, we similarly find
3753 \begin{tabular}{ccc}
3754 % e0
3755 \verb+{0,1,2}+ & $(0,0,0)$ & $p \ge 0$, $-q + r \ge 0$ and $q \ge 0$
3757 % 70
3758 \verb+{1,2,3}+ & $(p,0,0)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3760 % c8
3761 \verb+{0,1,4}+ & $(0,0,-p+r)$ & $-q + r \ge 0$, $p \ge 0$ and $q \ge 0$
3763 % 58
3764 \verb+{1,3,4}+ & $(p,0,r)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3766 % a4
3767 \verb+{0,2,5}+ & $(0,q,0)$ & $q \ge 0$, $p \ge 0$ and $-q + r \ge 0$
3769 % 34
3770 \verb+{2,3,5}+ & $(p, 2p+q, 0)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3772 % 8c
3773 \verb+{0,4,5}+ & $(0, q, -p+r)$ & $q \ge 0$, $-q + r \ge 0$ and $p \ge 0$
3775 % 1c
3776 \verb+{3,4,5}+ & $(p, 2p+q, r)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3777 \end{tabular}
3779 Combining these constraints with the initial constraints of the problem
3780 on the parameters
3781 $p \ge 0$, $q \ge 0$ and $r \ge 0$, we find the chamber
3783 \{\,
3784 (p,q,r) \mid p \ge 0 \wedge -p + r \ge 0 \wedge q \ge 0
3785 \,\}
3787 For the second chamber, we have
3789 \begin{tabular}{ccc}
3790 % 70
3791 \verb+{1,2,3}+ & $(p,0,0)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3793 % 58
3794 \verb+{1,3,4}+ & $(p,0,r)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3796 % 34
3797 \verb+{2,3,5}+ & $(p, 2p+q, 0)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3799 % 1c
3800 \verb+{3,4,5}+ & $(p, 2p+q, r)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3802 % 64
3803 \verb+{1,2,5}+ & $(-\frac q 2,0,0)$ &
3804 $-q \ge 0$, $2p + q \ge 0$ and $-2p -q+2r \ge 0$
3806 % 4c
3807 \verb+{1,4,5}+ & $(-\frac q 2,0,-p-\frac q 2+r)$ &
3808 $-q \ge 0$, $-2p -q+2r \ge 0$ and $2p + q \ge 0$
3809 \end{tabular}
3811 The chamber is therefore
3813 \{\,
3814 (p,q,r) \mid q = 0 \wedge p \ge 0 \wedge -p +r \ge 0
3815 \,\}
3817 Note that by intersecting with the initial constraints this chamber
3818 is no longer full-dimensional and can therefore be discarded.
3819 Finally, for the third chamber, we have
3821 \begin{tabular}{ccc}
3822 % 70
3823 \verb+{1,2,3}+ & $(p,0,0)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3825 % 58
3826 \verb+{1,3,4}+ & $(p,0,r)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3828 % 34
3829 \verb+{2,3,5}+ & $(p, 2p+q, 0)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3831 % 1c
3832 \verb+{3,4,5}+ & $(p, 2p+q, r)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3834 % 68
3835 \verb+{1,2,4}+ & $(p-r,0,0)$ &
3836 $p -r \ge 0$, $r \ge 0$ and $2p +q -2r \ge 0$
3838 % 2c
3839 \verb+{2,4,5}+ & $(p-r,2p+q-2r, 0)$ &
3840 $p -r \ge 0$, $2p +q -2r \ge 0$ and $r \ge 0$
3841 \end{tabular}
3843 The chamber is therefore
3845 \{\,
3846 (p,q,r) \mid p - r \ge 0 \wedge q \ge 0 \wedge r \ge 0
3847 \,\}
3849 \end{example}
3851 Now let us consider general parametric polytopes.
3852 First note that we can follow the same procedure as above
3853 if we replace $\vec x$ by $\vec x' - \vec c(\vec p)$
3854 in \eqref{eq:TOPCOM:polytope}, i.e.,
3855 if our problem has the form
3856 \begin{equation}
3857 \label{eq:TOPCOM:polytope:2}
3858 \begin{cases}
3859 \begin{aligned}
3860 \vec x' &\ge \vec c(\vec p)
3862 A \, \vec x' &\le \vec b(\vec p) + A \vec c(\vec p)
3864 \end{aligned}
3865 \end{cases}
3866 \end{equation}
3867 as saturating a constraint $x_i = 0$ is equivalent
3868 to saturating the constraint $x_i' = c_i(\vec p)$
3869 and, similarly, $\sps {\vec a_j}{\vec x} = b_j(\vec p)$
3870 is equivalent to
3871 $\sps {\vec a_j}{\vec x'} = b_j(\vec p) + \sps {\vec a_j}{\vec c(\vec p)}$.
3873 In the general case, the problem has the form
3875 A \vec x \ge \vec b(\vec p)
3877 and then we apply the technique of \autoref{s:standard}.
3878 Let $A'$ be a non-singular square submatrix of $A$ with the same number
3879 of columns and compute the (left) \indac{HNF} $H = A' U$ with $U$ unimodular
3880 and $H$ lower-triangular with non-positive elements below the diagonal.
3881 Replacing $\vec x$ by $U \vec x'$, we obtain
3883 \begin{cases}
3884 \begin{aligned}
3885 H \vec x' &\ge \vec b'(\vec p)
3887 -A''U \, \vec x' &\le -\vec b''(\vec p)
3889 \end{aligned}
3890 \end{cases}
3892 with $A''$ the remaining rows of $A$ and $\vec b(\vec p)$ split
3893 in the same way.
3894 If $H$ happens to be the identity matrix, then our problem is
3895 of the form \eqref{eq:TOPCOM:polytope:2} and we already know how
3896 to solve this problem.
3897 Note that, again, saturating any of the transformed constraints
3898 in $\vec x'$ is equivalent to saturating the corresponding constraint
3899 in $\vec x$. We therefore only need to compute $-A'' U$ for the
3900 computation of the kernel $K$. To construct the parametric vertices
3901 in the original coordinate system, we can simply use the original
3902 constraints.
3903 The same reasoning holds if $H$ is any diagonal matrix, since
3904 we can virtually replace $H \vec x$ by $\vec x'$ without affecting
3905 the non-negativity of the variables.
3907 If $H$ is not diagonal, then we can introduce new constraints
3908 $x'_j \ge d(\vec p)$, where $d(\vec p)$ is some symbolic constant.
3909 These constraints do not remove any solutions
3910 since each row in $H$ expresses that the corresponding variable is
3911 greater than or equal to a non-negative combination of the
3912 previous variables plus some constant.
3913 We can then proceed as before. However, to reduce unnecessary computations
3914 we may remove from $K$ the rows that correspond to these new rows.
3915 Any solution saturating the new constraint, would also saturate
3916 the corresponding constraint $\vec h_j^\T$ and all
3917 the constraints corresponding to the non-zero
3918 entries in $\vec h_j^\T$.
3919 If a chamber contains a vertex obtained by saturating such a new
3920 constraint, it would appear multiple times in the same chamber,
3921 each time combined with different constraints from the above set.
3922 Furthermore, there would also be another (as it turns out, identical)
3923 chamber where the vertex is only defined by the other constraints.
3925 \begin{example}
3926 Consider the parametric polytope
3928 P(n) = \{\,
3929 (i,j) \mid
3930 1 \le i \wedge 2 i \le 3 j \wedge j \le n
3931 \,\}
3934 The constraints are
3936 \begin{bmatrix}
3937 1 & 0 \\
3938 -2 & 3 \\
3939 0 & -1
3940 \end{bmatrix}
3941 \begin{bmatrix}
3942 i \\ j
3943 \end{bmatrix}
3945 \begin{bmatrix}
3946 1 \\
3947 0 \\
3949 \end{bmatrix}
3952 The top $2 \times 2$ submatrix is already in \indac{HNF}.
3953 We have $3 j \ge 2i \ge 2$, so we can add a constraint
3954 of the form $j \ge c(n)$ and obtain
3956 \begin{bmatrix}
3957 A & I
3958 \end{bmatrix}
3960 \begin{bmatrix}
3961 0 & 1 & 1 & 0
3963 2 & -3 & 0 & 1
3964 \end{bmatrix}
3967 while $K$ with $\begin{bmatrix}A & I\end{bmatrix} K = 0$ is given
3970 \begin{bmatrix}
3971 0 & 1 & 1 & 0
3973 2 & -3 & 0 & 1
3974 \end{bmatrix}
3975 \begin{bmatrix}
3976 1 & 0 \\
3977 0 & 1 \\
3978 0 & -1 \\
3979 -2 & 3
3980 \end{bmatrix}
3983 The second row of $K$ corresponds to the second variable,
3984 which in turn corresponds to the newly added constraint.
3985 Passing all rows of $K$ to \ai[\tt]{TOPCOM} we would get
3986 \begin{verbatim}
3987 > points2triangs --regular <<EOF
3988 > [[1 0],[0,1],[0,-1],[-2,3]]
3989 > EOF
3990 T[1]:={{0,1},{0,2},{1,3},{2,3}};
3991 T[2]:={{0,2},{2,3},{0,3}};
3992 T[3]:={};
3993 \end{verbatim}
3994 The first vertex in the first chamber saturates the second row
3995 (row 1) and therefore saturates both the first (0) and fourth (3)
3996 and it appears a second time as \verb+{1,3}+. Combining
3997 these ``two'' vertices into one as \verb+{0,3}+ results in the
3998 second (identical) chamber.
3999 Removing the row corresponding to the new constraint from $K$
4000 we remove the duplicates
4001 \begin{verbatim}
4002 > points2triangs --regular <<EOF
4003 > [[1 0],[0,-1],[-2,3]]
4004 > EOF
4005 T[1]:={{0,1},{1,2},{0,2}};
4006 T[2]:={};
4007 \end{verbatim}
4008 Note that in this example, we also could have interchanged
4009 the second and the third constraint and then have replaced $j$ by $-j'$.
4010 \end{example}
4012 In practice, this method of computing a \ai{chamber decomposition}
4013 does not seem to perform very well, mostly because
4014 \ai[\tt]{TOPCOM} can not exploit all available information
4015 about the parametric polytopes and will therefore compute
4016 many redundant triangulations/chambers.
4017 In particular, any chamber that does not intersect with
4018 the parameter domain of the parametric polytope, or only
4019 intersects in a face of this parameter domain, is completely redundant.
4020 Furthermore, if the parametric polytope is not simple, then many
4021 different combinations of the constraints will lead to the same parametric
4022 vertex. Many triangulations will therefore correspond to one and the
4023 same chamber in the chamber complex of the parametric polytope.
4024 For example, for a dilated octahedron, \ai[\tt]{TOPCOM} will
4025 compute 150 triangulations/chambers, 104 of which are empty,
4026 while the remaining 46 refer to the same single chamber.
4029 \subsection{Computing the Hilbert basis of a cone}
4030 \label{s:hilbert}
4032 To compute the \ai{Hilbert basis} of a cone, we use
4033 the \ai[\tt]{zsolve} library from \ai[\tt]{4ti2} \shortcite{4ti2},
4034 which implements the technique of \shortciteN{Hemmecke2002Hilbert}.
4035 We first remove all equalities from the cone through unimodular
4036 transformations and then apply the technique of \autoref{s:standard}
4037 to put the cone in ``standard form''. Note that for a (non-parametric)
4038 cone the constant term $\vec b$ in \eqref{eq:non-standard} is $\vec 0$.
4039 The constraints $D \vec x \ge \vec c = \vec 0$ of \eqref{eq:standard:2}
4040 are therefore equivalent to $\vec x \ge \vec 0$.
4043 \subsection{Integer Feasibility}
4044 \label{s:feasibility}
4046 For testing whether a polytope $P \subset \QQ^d$ contains any integer points,
4047 we use the technique of~\shortciteN{Cook1993implementation},
4048 based on \ai{generalized basis reduction}.
4050 The technique basically looks for a ``short vector'' $\vec c$ in the
4051 lattice $\ZZ^d$, where shortness is measured in terms of
4052 the \ai{width} of the polytope $P$ along that direction,
4053 \begin{align*}
4054 \width_{\vec c} P
4056 \max \{\, \sp c x \mid \vec x \in P \,\}
4058 \min \{\, \sp c x \mid \vec x \in P \,\}
4061 \max \{\, \sps {\vec c} {\vec x - \vec y} \mid \vec x, \vec y \in P \,\}
4063 \end{align*}
4064 The \defindex{lattice width} is the minimum width over all
4065 non-zero integer directions:
4067 \width P =
4068 \min_{\vec c \in \ZZ^d \setminus \{ \vec 0 \} } \width_{\vec c} P
4071 If the dimension $d$ is fixed then
4072 the lattice width of any polytope $P \subset \QQ^d$
4073 containing no integer points is bounded by a constant%
4074 ~\shortcite{Lagarias90,Barvinok02,Banaszczyk1999flatness}.
4075 If we slice the polytope using hyperplanes orthogonal
4076 to a short direction, i.e., a direction where the width
4077 is small, we will therefore only need to inspect
4078 ``few'' of them before either finding one with an integer point,
4079 or running out of hyperplanes, meaning that the
4080 polytope did not contain any integer points.
4081 Each slice is checked for integer points by applying
4082 the above method recursively.
4084 A nice feature of this technique is that it will
4085 not only tell you if there is any integer point
4086 in the given polytope, but it will actually compute
4087 one if there is any.
4089 The short vector is obtained as the first vector
4090 of a ``reduced basis'' of the lattice $\ZZ^d$ with respect
4091 to the polytope.
4092 In particular, the first vector $\vec b_1$ of this reduced basis
4093 will satisfy
4095 \width_{\vec b_1} P
4097 \frac{\width P}
4098 {\left(\frac 1 2 - \varepsilon\right)^{d-1}}
4101 with $0 < \varepsilon < 1/2$ a fixed constant.
4102 That is, the width in direction $\vec b_1$ is no more than a constant
4103 factor bigger than the lattice width.
4104 See~\shortcite{Cook1993implementation} for details.
4105 In our implementation we use $\varepsilon = 1/4$.
4106 When used in the above integer feasibility testing algorithm,
4107 we will also terminate the reduced basis computation
4108 as soon as the width along the first basis vector is smaller than 2.
4109 This means that there will be at most 2 slices orthogonal to the chosen
4110 direction.
4112 The computation of the above reduced basis requires the solution
4113 of many linear programs, for which we use any of the following
4114 external solvers:
4115 \begin{itemize}
4116 \item \ai[\tt]{GLPK}~\shortcite{GLPK}
4118 This solver is based on double precision floating point arithmetic and
4119 may therefore not be suitable if the coefficients of the constraints
4120 describing the polytope are large.
4122 \item \ai[\tt]{cdd}~\shortcite{cdd}
4124 This solver is based on exact integer arithmetic.
4125 Note that you need version \verb+cddlib 0.94e+ or newer.
4126 Earlier versions (\verb+0.93+--\verb+0.94d+) have
4127 a bug that may sometimes result in a polytope being
4128 reported as (rationally) empty even though it is not.
4130 \item \piplib/~\shortcite{Feautrier:PIP}
4132 This solver is also based on exact integer arithmetic
4133 and uses the \ai{dual simplex} method to solve a linear program.
4134 Two versions are available, \ai[\tt]{pip} will present the
4135 original program to \piplib/, while \ai[\tt]{pip-dual} will present
4136 the dual program to \piplib/, effectively having it apply the primal
4137 simplex method to the original problem.
4138 The latter may seem more appropriate since the computation
4139 of the reduced basis only requires the dual solution of
4140 any linear program. However, in practice, it appears
4141 that \ai[\tt]{pip} is often faster than \ai[\tt]{pip-dual}.
4142 \end{itemize}
4143 The LP solver to use can be selected with the \ai[\tt]{--gbr} option.
4146 \subsection{Computing the integer hull of a polyhedron}
4147 \label{s:integer:hull}
4149 For computing the \ai{integer hull} of a polyhedron,
4150 we first describe how to compute the convex hull of a set
4151 given as an oracle for optimizing a linear objective
4152 function over the set and then
4153 we explain how to optimize a linear objective function over
4154 the integer points of a polyhedron.
4155 Applying the first with the second as \ai{optimization oracle}
4156 yields a method for computing the requested integer hull.
4158 \subsubsection{Computing the convex hull based on an optimization oracle}
4160 The algorithm described below is presented by
4161 \shortciteN[Remark~2.5]{Cook1992} as an extension of the
4162 algorithm by \shortciteN[Section~3]{Edmonds82} for computing
4163 the {\em dimension} of a polytope for which only an optimization oracle
4164 is available. The algorithm is described in a bit more detail
4165 by \shortciteN{Eisenbrand2000PhD} and reportedly stems from
4166 \shortciteN{Hartmann1989PhD}.
4167 Essentially the same algorithm has also been implemented
4168 by \shortciteN{Huggins06}, citing
4169 \ai{beneath/beyond}~\shortcite{Preparata1985} as his inspiration.
4171 The algorithm start out from an initial set of points from
4172 the set $S$. After computing the convex hull of this set
4173 of points, we take one of its bounding constraints and use
4174 the optimization oracle
4175 to compute an optimal point in $S$ (but on the other side
4176 of the bounding hyperplane) along the
4177 outer normal of this bounding constraint.
4178 If a new point is found, it is added to the set of points
4179 and a new convex hull is computed, or the old one is adapted
4180 in a beneath/beyond fashion. Otherwise, the chosen bounding constraint
4181 is also a bounding constraint of $S$ and need not be considered anymore.
4182 The process continues until all bounding constraints in the
4183 description of the current convex hull have been considered.
4185 In principle, the initial set of points in the above algorithm
4186 may be empty, with a ``convex hull'' described by a set of
4187 conflicting constraints and each equality in the description of any
4188 intermediate lower-dimensional convex hull being considered
4189 as a pair of bounding constraints with opposite outer normals.
4190 However, in our implementation, we have chosen to first compute
4191 a maximal set of affinely independent points by first taking any
4192 point from $S$ and then adding points from $S$ not on one of
4193 the equalities satisfied by all points found so far.
4194 This allows us to not have to worry about equalities in the
4195 main algorithm.
4196 In the case of the computation of the integer hull, finding
4197 these affinely independent points can be accomplished using the technique of
4198 \autoref{s:feasibility}.
4200 \begin{figure}
4201 \intercol=0.58cm
4202 \begin{xy}
4203 <\intercol,0pt>:<0pt,\intercol>::
4204 \POS(-1,0)*\xybox{
4205 \def\latticebody{\POS="c"+(0,0.5)\ar@{--}"c"+(0,6.5)}%
4206 \POS0,{\xylattice{1}{6}00}%
4207 \def\latticebody{\POS="c"+(0.5,0)\ar@{--}"c"+(6.5,0)}%
4208 \POS0,{\xylattice00{1}6}%
4209 \POS@i@={(1.5,2.75),(5.75,2.25),(5.5,5.25),(2.75,4.75),(1.5,2.75)},
4210 {0*\xypolyline{}}
4211 \POS@i@={(2,3),(3,3),(3,4),(2,3)},{0*[|(3)]\xypolyline{}}
4212 \POS(2,3)*{\bullet}
4213 \POS(3,3)*{\bullet}
4214 \POS(3,4)*{\bullet}
4215 \POS(3,3.5)\ar(3.5,3.5)
4216 \POS(5,3)*{\circ}
4217 \POS(5,4)*{\circ}
4218 \POS(5,5)*{\circ}
4220 \POS(6,0)*\xybox{
4221 \def\latticebody{\POS="c"+(0,0.5)\ar@{--}"c"+(0,6.5)}%
4222 \POS0,{\xylattice{1}{6}00}%
4223 \def\latticebody{\POS="c"+(0.5,0)\ar@{--}"c"+(6.5,0)}%
4224 \POS0,{\xylattice00{1}6}%
4225 \POS@i@={(1.5,2.75),(5.75,2.25),(5.5,5.25),(2.75,4.75),(1.5,2.75)},
4226 {0*\xypolyline{}}
4227 \POS@i@={(2,3),(5,3),(3,4),(2,3)},{0*[|(3)]\xypolyline{}}
4228 \POS(2,3)*{\bullet}
4229 \POS(5,3)*{\bullet}
4230 \POS(3,4)*{\bullet}
4231 \POS(4,3.5)\ar(4.25,4)
4232 \POS(5,5)*{\circ}
4234 \POS(13,0)*\xybox{
4235 \def\latticebody{\POS="c"+(0,0.5)\ar@{--}"c"+(0,6.5)}%
4236 \POS0,{\xylattice{1}{6}00}%
4237 \def\latticebody{\POS="c"+(0.5,0)\ar@{--}"c"+(6.5,0)}%
4238 \POS0,{\xylattice00{1}6}%
4239 \POS@i@={(1.5,2.75),(5.75,2.25),(5.5,5.25),(2.75,4.75),(1.5,2.75)},
4240 {0*\xypolyline{}}
4241 \POS@i@={(2,3),(5,3),(5,5),(3,4),(2,3)},{0*[|(3)]\xypolyline{}}
4242 \POS(2,3)*{\bullet}
4243 \POS(5,3)*{\bullet}
4244 \POS(5,5)*{\bullet}
4245 \POS(3,4)*{\bullet}
4247 \end{xy}
4248 \caption{The integer hull of a polytope}
4249 \label{f:integer:hull}
4250 \end{figure}
4252 \begin{example}
4253 Assume we want to compute the integer hull of the polytope in the left part
4254 of \autoref{f:integer:hull}.
4255 We first compute a set of three affinely independent points,
4256 shown in the same part of the figure.
4257 Of the three facets of the corresponding convex hull,
4258 optimization along the outer normal (depicted by an arrow in the figure)
4259 of only one facet will yield any additional points. The other two
4260 are therefore facets of the integer hull.
4261 Optimization along the above outer normal may yield any of the
4262 points marked by a $\circ$.
4263 Assuming it is the bottom one, we end up with the updated
4264 convex hull in the middle of the figure. This convex hull
4265 has only one new facet. Adding the point found by optimizing
4266 over this facet's outer normal, we obtain the convex hull
4267 on the right of the figure.
4268 There are two new facets, but neither of them yields any
4269 further points. We have therefore found the integer hull
4270 of the polytope.
4271 \end{example}
4273 \subsubsection{Optimization over the integer points of a polyhedron}
4274 \label{s:optimization}
4276 We assume that we want to find the {\em minimum} of
4277 some linear objective function. When used in the computation
4278 of the integer hull of some polytope, the objective function
4279 will therefore correspond to the inner normal of some facet.
4281 During our search for an optimal integer point with respect
4282 to some objective function, we will keep track of the best
4283 point so far as well as a lower bound $l$
4284 and an upper bound $u$ such that the value at the optimal point
4285 (if it is better than the current best) lies between those
4286 two bounds.
4287 Initially, there is no best point yet and values for $l$ and $u$
4288 may be obtained from optimization over the linear relaxation.
4289 When used in the computation of the integer hull of some polytope,
4290 the upper bound $u$ is one less than the value attained on
4291 the given facet of the current approximation.
4293 As long as $l \le u$, we perform the following steps
4294 \begin{itemize}
4295 \item use the integer feasibility technique of \autoref{s:feasibility}
4296 to test whether there is any integer point with value in
4297 $[l,u']$, where $u'$ is
4298 \begin{itemize}
4299 \item $u$ if the previous test for an integer point did not produce a point
4300 \item $l+\floor{\frac{u-l-1}2}$
4301 if the previous test for an integer point {\em did\/} produce a point
4302 \end{itemize}
4303 \item if a point is found, then remember it as the current best
4304 and replace $u$ by the value at this point minus one,
4305 \item otherwise, replace $l$ by $u'+1$.
4306 \end{itemize}
4307 When used in the computation of the integer hull of some polytope,
4308 it is useful to not only keep track of the best point so far,
4309 but of all points found.
4310 These points will all lie outside of the current approximation
4311 of the integer hull and adding them all instead of just one,
4312 will typically get us to the complete integer hull quicker.
4314 \begin{figure}
4315 \intercol=0.7cm
4316 \begin{xy}
4317 <\intercol,0pt>:<0pt,\intercol>::
4318 \POS(0.5,0)\ar@{-}(16.5,0)
4319 \def\latticebody{\POS="c"+(0,-0.2)\ar@{--}"c"+(0,0.2)\POS"c"*++!D{\the\latticeA}}%
4320 \POS0,{\xylattice{1}{16}00}%
4321 \POS(6,0)*!C{\bullet}
4322 \POS(7,0)*{\bullet}
4323 \POS(8,0)*{\bullet}
4324 \POS(12,0)*{\bullet}
4325 \POS(13,0)*{\bullet}
4326 \POS(14,0)*{\bullet}
4327 \POS(15,0)*{\bullet}
4328 \POS(16,0)*{\bullet}
4329 \POS(1,-1)\ar@{-}(16,-1)
4330 \POS(8,-1)*{\bullet}
4331 \POS(1,-2)\ar@{-}(4,-2)
4332 \POS(5,-3)\ar@{-}(7,-3)
4333 \POS(6,-3)*{\bullet}
4334 \POS(4.9,-4)\ar@{-}(5.1,-4)
4335 \end{xy}
4336 \caption{The integer points of a polytope projected on an objective function}
4337 \label{f:hull:projected}
4338 \end{figure}
4340 \begin{example}
4341 \label{ex:hull:projected}
4342 Assume that the values of some objective function attained
4343 by the integer points of some polytope are as shown in
4344 \autoref{f:hull:projected} and assume we know that the optimal
4345 value lies between 1 and 16.
4346 In the first step we would look for a point attaining a value
4347 in the interval $[1,16]$. Suppose this yields a point attaining
4348 the value $8$ (second line of the figure). We record this point
4349 as the current best and update the search interval to $[1,7]$.
4350 In the second step, we look for a point attaining a value
4351 in the interval $[1,4]$, but find nothing and set the search interval
4352 to $[5,7]$.
4353 In the third step, we consider the interval $[5,7]$ and find
4354 a point attaining the value 6. We update the current best value
4355 and set the search interval to $[5,5]$.
4356 In the fourth step, we consider the interval $[5,5]$, find no
4357 points and update the interval to ``$[6,5]$''.
4358 Since the lower bound is now larger than the upper bound, the
4359 algorithm terminates, returning the best or all point(s) found.
4360 \end{example}
4363 \subsection{Computing the integer hull of a truncated cone}
4364 \label{s:hull:cone}
4366 In \autoref{s:width} we will need to compute the \ai{integer hull}
4367 of a cone with the origin removed ($C \setminus \{ \vec 0 \}$).
4369 \subsubsection{Using the Hilbert basis of the cone}
4371 As proposed by \shortciteN{Koeppe2007personal},
4372 one way of computing this integer hull is to first compute
4373 the \ai{Hilbert basis} of $C$ (see \autoref{s:hilbert})
4374 and to then remove from that Hilbert basis the points that
4375 are not vertices of the integer hull of $C \setminus \{ \vec 0 \}$.
4376 The Hilbert basis of $C$ is the minimal set of points
4377 $\vec b_i \in C \cap \ZZ^d$ such that every integer point
4378 $\vec x \in C \cap \ZZ^d$ can be written as a non-negative
4379 {\em integer} combination of the $\vec b_i$.
4380 The vertices $\vec v_j$ of the integer hull of $C \setminus \{ \vec 0 \}$
4381 are such that every integer point
4382 $\vec x \in (C \cap \ZZ^d) \setminus \{ \vec 0 \}$ can
4383 be written as s non-negative {\em rational} combination of $\vec v_j$.
4384 Clearly, any $\vec v_j$ is also a $\vec b_i$ since $\vec v_j$ can
4385 not be written as the sum of a (rational) convex combination of
4386 other integer points in $(C \cap \ZZ^d) \setminus \{ \vec 0 \}$
4387 and a non-negative combination of the extremal rays $\vec r_k$ of $C$.
4388 A fortiori, it can therefore not be written as an integer combination
4389 of other integer points in $C$.
4390 To obtain the $\vec v_j$ from the $\vec b_i$ we therefore simply
4391 need to remove first $(0,0)$ and then those $\vec b_i$ that are
4392 not an extremal ray and that {\em can} be written as a combination
4394 \vec b_i = \sum_{j \ne i} \vec \alpha_j \vec b_j + \sum_k \beta_k \vec r_k
4395 \qquad\text{with $\alpha_j, \beta_k \ge 0$ and $\sum_{j \ne i} \alpha_j = 1$}
4398 Since the $\vec r_k$ are also among the $\vec b_j$, this can
4399 be simplified to checking whether there exists a rational
4400 solution for $\vec \alpha_j$ to
4402 \vec b_i = \sum_{j \ne i} \vec \alpha_j \vec b_j
4403 \qquad\text{with $\alpha_j \ge 0$ and $\sum_{j \ne i} \alpha_j \ge 1$}
4407 \begin{figure}
4408 \intercol=1.1cm
4409 \begin{xy}
4410 <\intercol,0pt>:<0pt,\intercol>::
4411 \POS@i@={(3,-4.5),(2,-3),(1,-1),(1,1),(3,4),(4.125,5.5),(5.5,5.5),(5.5,-4.5)},{0*[grey]\xypolyline{*}}
4412 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4413 \POS0,{\xylattice{-0}{5}00}%
4414 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4415 \POS0,{\xylattice00{-4}5}%
4416 \POS0\ar(2,-3)
4417 \POS0\ar(3,4)
4418 \POS(2,-3)*{\bullet}
4419 \POS(3,4)*{\bullet}
4420 \POS(1,1)*{\bullet}
4421 \POS(1,-1)*{\bullet}
4422 \POS(1,0)*{\bullet}
4423 \POS(2,-3)*{\times}
4424 \POS(3,4)*{\times}
4425 \POS(1,1)*{\times}
4426 \POS(1,-1)*{\times}
4427 \end{xy}
4428 \caption{The Hilbert basis and the integer hull of a truncated cone}
4429 \label{f:hilbert:hull}
4430 \end{figure}
4432 \begin{example} \label{ex:hilbert:hull}
4433 Consider the cone
4435 C = \poshull \,\{(2,-3), (3,4)\}
4438 shown in Figure~\ref{f:hilbert:hull}.
4439 The Hilbert basis of this cone is
4440 $$\{(0,0),(2,-3),(3,4),(1,1),(1,-1),(1,0)\}.$$
4441 We have $(1,0) = \frac 1 2 (1,1) + \frac 1 2 (1,-1)$,
4442 while $(1,1)$ and $(1,-1)$ can not be written as
4443 overconvex combinations of the other $\vec b_i \ne \vec 0$.
4444 The vertices of the integer hull of $C \setminus \{ \vec 0 \}$
4445 are therefore
4446 $$\{(2,-3),(3,4),(1,1),(1,-1)\}.$$
4447 \end{example}
4449 \subsubsection{Using generalized basis reduction}
4450 \label{s:hull:cone:gbr}
4452 Another way of computing the integer hull of a truncated cone is to apply
4453 the method of \autoref{s:integer:hull}.
4454 In this case, the initial set of points will consist
4455 of (the smallest integer representatives of) the extremal rays
4456 of the cone, together with the extremal rays themselves.
4457 That is, if $C = \poshull \, \{ \vec r_j \}$ with
4458 $\vec r_j \in \ZZ^d$, then our initial approximation of the
4459 integer hull of $C \setminus \{ \vec 0 \}$ is
4461 \convhull \, \{ \vec r_j \} + \poshull \, \{ \vec r_j \}
4464 Furthermore, we need never consider any
4465 of the bounding constraints that are also bounding constraints
4466 of the original cone.
4467 When optimizing along the normal of any of the other facets, we can
4468 take the lower bound to be $1$. This will ensure that
4469 the origin is excluded, without excluding any other integer points.
4471 \begin{figure}
4472 \intercol=0.5cm
4473 \begin{xy}
4474 <\intercol,0pt>:<0pt,\intercol>::
4475 \POS(0,0)*\xybox{
4476 \POS@i@={(3,-4.5),(2,-3),(3,4),(4.125,5.5),(5.5,5.5),(5.5,-4.5)},{0*[grey]\xypolyline{*}}
4477 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4478 \POS0,{\xylattice{-0}{5}00}%
4479 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4480 \POS0,{\xylattice00{-4}5}%
4481 \POS0\ar(2,-3)
4482 \POS0\ar(3,4)
4483 \POS(2,-3)*{\bullet}
4484 \POS(3,4)*{\bullet}
4485 \POS(1,1)*{\circ}
4487 \POS(8,0)*\xybox{
4488 \POS@i@={(3,-4.5),(2,-3),(1,1),(3,4),(4.125,5.5),(5.5,5.5),(5.5,-4.5)},{0*[grey]\xypolyline{*}}
4489 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4490 \POS0,{\xylattice{-0}{5}00}%
4491 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4492 \POS0,{\xylattice00{-4}5}%
4493 \POS0\ar(2,-3)
4494 \POS0\ar(3,4)
4495 \POS(2,-3)*{\bullet}
4496 \POS(3,4)*{\bullet}
4497 \POS(1,1)*{\bullet}
4498 \POS(1,-1)*{\circ}
4500 \POS(16,0)*\xybox{
4501 \POS@i@={(3,-4.5),(2,-3),(1,-1),(1,1),(3,4),(4.125,5.5),(5.5,5.5),(5.5,-4.5)},{0*[grey]\xypolyline{*}}
4502 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4503 \POS0,{\xylattice{-0}{5}00}%
4504 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4505 \POS0,{\xylattice00{-4}5}%
4506 \POS0\ar(2,-3)
4507 \POS0\ar(3,4)
4508 \POS(2,-3)*{\bullet}
4509 \POS(3,4)*{\bullet}
4510 \POS(1,1)*{\bullet}
4511 \POS(1,-1)*{\bullet}
4513 \end{xy}
4514 \caption{The integer hull of a truncated cone}
4515 \label{f:cone:integer:hull}
4516 \end{figure}
4518 \begin{example}
4519 Consider once more the cone
4521 C = \poshull \,\{(2,-3), (3,4)\}
4523 from Example~\ref{ex:hilbert:hull}.
4524 The initial approximation is
4526 C = \convhull \,\{(2,-3), (3,4)\} + \poshull \,\{(2,-3), (3,4)\}
4529 which is shown on the left of \autoref{f:cone:integer:hull}.
4530 The only bounding constraint that does not correspond to a
4531 bounding constraint of $C$ is $7 x - y \ge 17$.
4532 In the first step, we will therefore look for a point
4533 minimizing $7 x - y$ with values in the interval $[1,16]$.
4534 All values of this objective function in the given interval
4535 attained by points in $C$ are shown in \autoref{f:hull:projected}.
4536 From Example~\ref{ex:hull:projected}, we know that the optimal
4537 value is $6$ and this corresponds to the point $(1,1)$.
4538 Adding this point to our hull, we obtain the approximation
4539 in the middle of \autoref{f:cone:integer:hull}.
4540 This approximation has two new facets.
4541 The bounding constraint $3x - 2 y \ge 1$ will not produce
4542 any new points since we would be looking for one in the
4543 interval ``$[1,0]$''.
4544 The other new bounding constraint is $4x + y \ge 5$.
4545 Minimizing $4 x+ y$ with values in the interval $[1,4]$,
4546 we find the minimal value $3$ corresponding to the point $(1,-1)$.
4547 Adding this point, we obtain the complete integer hull
4548 shown on the right of \autoref{f:cone:integer:hull}.
4549 Note that if in the first step we would have added not only
4550 the point corresponding to the optimal value, but instead
4551 all points found in Example~\ref{ex:hull:projected},
4552 then we would have obtained the complete integer hull directly.
4553 \end{example}
4556 \subsection{Computing the lattice width of a parametric polytope}
4557 \label{s:width}
4559 To compute the \ai{lattice width} of a \ai{parametric polytope},
4560 we essentially use the technique of \shortciteN{Eisenbrand2007parameterised},
4561 which improves upon the technique of \shortciteN{Kannan1992}.
4562 Given a parametric polytope
4564 P(\vec p) = \{\, \vec x \mid A \vec x + \vec b(\vec p) \ge \vec 0 \,\}
4567 the width along a direction $\vec c$ is defined in the same
4568 way as for non-parametric polytopes (see \autoref{s:feasibility}),
4569 \begin{equation}
4570 \label{eq:width}
4571 \width_{\vec c} P(\vec p)
4573 \max \{\, \sp c x \mid \vec x \in P(\vec p) \,\}
4575 \min \{\, \sp c x \mid \vec x \in P(\vec p) \,\}
4577 \end{equation}
4578 The \defindex{lattice width} is the minimum width over all
4579 non-zero integer directions:
4581 \width P(\vec p) =
4582 \min_{\vec c \in \ZZ^d \setminus \{ \vec 0 \} } \width_{\vec c} P(\vec p)
4585 We assume that the parameter domain $Q$ of $P(\vec p)$, i.e., the
4586 set of parameter values for which $P(\vec p) \ne \emptyset$,
4587 is full-dimensional and that for each $\vec p$ from the interior
4588 of $Q$, $P(\vec p)$ is also full-dimensional.
4590 Clearly, for any given direction $\vec c$, the minimum and
4591 maximum in \eqref{eq:width} are attained at (different)
4592 vertices of $P(\vec p)$.
4593 The idea of the algorithm is then to consider all pairs
4594 of parametric vertices of $P(\vec p)$, to compute all candidate
4595 integer directions for a given pair of vertices and then to
4596 compute the minimum width over all candidate integer directions
4597 found.
4599 For any given parametric vertex $\vec v(\vec p)$, the (rational)
4600 directions for which this vertex is minimal can be found as follows.
4601 Let $\vec v(\vec p) + C$ be the \ai{vertex cone} of $\vec v(\vec p)$.
4602 If $\vec v(\vec p)$ is minimal for $\vec c$, then all other points
4603 in the vertex cone must yield a bigger or equal value, i.e.,
4604 $\sp y c \ge 0$ for all $\vec y \in C$.
4605 The set of directions is therefore the \ai{polar cone} $C^*$ of $C$.
4606 Note that, in principle, we should only do this for pairs
4607 of vertices that have a common activity domain, where the
4608 activity domains have been partially opened using the
4609 technique of \autoref{p:inclusion-exclusion} to avoid
4610 multiple vertices that coincide on a lower-dimensional
4611 chamber to all be considered on this intersection.
4612 However, this optimization has currently not been implemented.
4614 Given a pair of vertices $\vec v_1(\vec p)$ and $\vec v_2(\vec p)$,
4615 we may assume that $\vec v_1(\vec p)$ attains the minimum and
4616 $\vec v_2(\vec p)$ attains the maximum.
4617 If $\vec v_1(\vec p) + C_1$ and $\vec v_2(\vec p) + C_2$ are the
4618 corresponding vertex cones, then the set of (rational) directions for this
4619 pair of vertices is
4621 C_{1,2} = \left( C_1^* \cap -C_2^* \right) \setminus \{ \vec 0 \}
4624 The set of candidate integer directions are therefore
4625 the vertices of the integer hull of $C_{1,2}$, which
4626 can be computed as explained in \autoref{s:hull:cone}.
4627 To see this, note that by construction
4628 $\sps {\vec c}{\vec v_1(\vec p)} \le \sps {\vec c}{\vec v_2(\vec p)}$
4629 and so
4631 w_{\vec c}(\vec p) = \width_{\vec c} P(\vec p)
4632 = \sps {\vec c}{\vec v_2(\vec p)-\vec v_1(\vec p)} \ge 0
4635 Any integer direction in $C_{1,2}$ will therefore yield
4636 a width that is at least as large as that of one
4637 of the vertices of the integer hull.
4638 Note that when using generalized basis reduction
4639 to compute the integer hull of these cones as in \autoref{s:hull:cone:gbr},
4640 it can be helpful to use as vertices for the initial approximation
4641 not only the extremal rays of the cone, but also those vertices
4642 of previously computed integer hulls that are elements of the current cone.
4644 After computing a list of all possible candidate width directions
4645 $\vec c_i$ and the corresponding widths $w_{\vec c_i}(\vec p)$,
4646 we keep only a single direction of all those that yield
4647 the same width (as an affine function of the parameters).
4648 Then we construct the chambers where each of the widths is minimal,
4649 i.e.,
4651 C_i = \{\, \vec p \in Q \mid \forall j :
4652 w_{\vec c_i}(\vec p) \le w_{\vec c_j}(\vec p) \,\}
4655 Note that many of the $C_i$ may be empty or of lower dimension
4656 than Q and that the other $C_i$ will intersect in common facets.
4657 To obtain a partition of partially-open full-dimensional chambers, we proceed
4658 as in \autoref{s:triangulation}.
4660 \begin{figure}
4661 \intercol=1.1cm
4662 \begin{xy}
4663 <\intercol,0pt>:<0pt,\intercol>::
4664 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
4665 \POS0,{\xylattice{-0}{10}00}%
4666 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(10.5,0)}%
4667 \POS0,{\xylattice00{-0}7}%
4668 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*[|(2)]\xypolyline{}}
4669 \POS(0,0)*{\bullet}
4670 \POS(5,3)*{\bullet}
4671 \POS(5,4)*{\bullet}
4672 \POS(9,6)*{\bullet}
4673 \POS(3,2)*{\bullet}
4674 \POS(4,3)*{\bullet}
4675 \POS(6,4)*{\bullet}
4676 \POS(7,5)*{\bullet}
4677 \POS(9,6);(8.7,6.4)**{}?(0)/1.1cm/="a"\POS(9,6)\ar"a"
4678 \POS(9,6);(9.1,5.8)**{}?(0)/1.1cm/="a"\POS(9,6)\ar"a"
4679 \POS(5,4);(5.4,3.5)**{}?(0)/1.1cm/="a"\POS(5,4)\ar"a"
4680 \POS(5,4);(5.1,3.8)**{}?(0)/1.1cm/="a"\POS(5,4)\ar"a"
4681 \POS(0,0);(0.4,-0.5)**{}?(0)/1.1cm/="a"\POS(0,0)\ar"a"
4682 \POS(0,0);(-0.3,0.5)**{}?(0)/1.1cm/="a"\POS(0,0)\ar"a"
4683 \POS(5,3);(4.7,3.5)**{}?(0)/1.1cm/="a"\POS(5,3)\ar"a"
4684 \POS(5,3);(4.7,3.4)**{}?(0)/1.1cm/="a"\POS(5,3)\ar"a"
4685 \POS(9,6)*+!DL{\vec v_1}
4686 \POS(0,0)*+!UR{\vec v_3}
4687 \POS(5,3)*+!UL{\vec v_4}
4688 \POS(5,4)*+!DR{\vec v_2}
4689 \end{xy}
4690 \caption{A polytope and its candidate width directions}
4691 \label{f:width}
4692 \end{figure}
4694 \begin{example} \label{ex:width}
4695 Consider the (non-parametric) polytope
4697 P = \left\{\,
4698 \vec x \mid
4699 \begin{aligned}
4700 -3 x_1 +5 x_2 &\ge 0 \\
4701 4 x_1 -5 x_2 &\ge 0 \\
4702 x_1 -2 x_2 + 3 &\ge 0 \\
4703 -3 x_1 +4 x_2 + 3 &\ge 0
4704 \end{aligned}
4705 \,\right\}
4707 shown in \autoref{f:width}. The polytope has four vertices
4709 \begin{aligned}
4710 \vec v_1 & = (9,6) \\
4711 \vec v_2 & = (5,4) \\
4712 \vec v_3 & = (0,0) \\
4713 \vec v_4 & = (5,3)
4715 \end{aligned}
4717 The corresponding cones of directions (for
4718 the given vertex to attain the minimum), also shown
4719 in \autoref{f:width} are
4721 \begin{aligned}
4722 C^*_1 & = \poshull \,\{ (-3,4), (1,-2) \} \\
4723 C^*_2 & = \poshull \,\{ (4,-5), (1,-2) \} \\
4724 C^*_3 & = \poshull \,\{ (4,-5), (-3,5) \} \\
4725 C^*_4 & = \poshull \,\{ (-3,5), (-3,4) \}
4727 \end{aligned}
4730 \begin{figure}
4731 \intercol=0.8cm
4732 \begin{xy}
4733 <\intercol,0pt>:<0pt,\intercol>::
4734 \def\latticebody{\POS="c"+(0,-6.5)\ar@{--}"c"+(0,2.5)}%
4735 \POS0,{\xylattice{-1}{5}00}%
4736 \def\latticebody{\POS="c"+(-1.5,0)\ar@{--}"c"+(5.5,0)}%
4737 \POS0,{\xylattice00{-6}2}%
4738 \POS0\ar@{->}(3,-4)\POS?!{(0,-6.5);(1,-6.5)}="a"
4739 \POS0\ar@{->}(-1,2)
4740 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4741 \POS0\ar@{->}(1,-2)
4742 \POS@i@={"a",(3,-4),(4,-5),"b"},{0*[grey]\xypolyline{*}}
4743 \POS0,{\ellipse(1.1)(*0;(4,3)*),^,(*0;(-2,-1)*){-}}
4744 \POS0,{\ellipse(1)(*0;(2,1)*),^,(*0;(5,4)*){-}}
4745 \POS0\ar@{->}(3,-4)\POS?!{(0,-6.5);(1,-6.5)}="a"
4746 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4747 \POS(4,-5)*{\bullet}
4748 \POS(3,-4)*{\bullet}
4749 \end{xy}
4750 \caption{The cone of directions $C_{2,1}$}
4751 \label{f:C:2:1}
4752 \end{figure}
4754 \begin{figure}
4755 \intercol=0.8cm
4756 \begin{xy}
4757 <\intercol,0pt>:<0pt,\intercol>::
4758 \def\latticebody{\POS="c"+(0,-6.5)\ar@{--}"c"+(0,5.5)}%
4759 \POS0,{\xylattice{-3}{5}00}%
4760 \def\latticebody{\POS="c"+(-3.5,0)\ar@{--}"c"+(5.5,0)}%
4761 \POS0,{\xylattice00{-6}5}%
4762 \POS0\ar@{->}(3,-4)
4763 \POS0\ar@{->}(-1,2)\POS?!{(0,5.5);(1,5.5)}="a"
4764 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4765 \POS0\ar@{->}(-3,5)
4766 \POS@i@={"b",(4,-5),(1,-1),(-1,2),"a",(5.5,5.5),(5.5,-6.5)},{0*[grey]\xypolyline{*}}
4767 \POS0\ar@{->}(-1,2)\POS?!{(0,5.5);(1,5.5)}="a"
4768 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4769 \POS0,{\ellipse(1.1)(*0;(4,3)*),^,(*0;(-2,-1)*){-}}
4770 \POS0,{\ellipse(1)(*0;(5,4)*),^,(*0;(-5,-3)*){-}}
4771 \POS(1,-1)*{\bullet}
4772 \POS(4,-5)*{\bullet}
4773 \POS(-1,2)*{\bullet}
4774 \end{xy}
4775 \caption{The cone of directions $C_{3,1}$}
4776 \label{f:C:3:1}
4777 \end{figure}
4779 \begin{figure}
4780 \intercol=0.8cm
4781 \begin{xy}
4782 <\intercol,0pt>:<0pt,\intercol>::
4783 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4784 \POS0,{\xylattice{-3}{3}00}%
4785 \def\latticebody{\POS="c"+(-3.5,0)\ar@{--}"c"+(3.5,0)}%
4786 \POS0,{\xylattice00{-4}5}%
4787 \POS0\ar@{->}(3,-4)
4788 \POS0\ar@{->}(-1,2)
4789 \POS0,{\ellipse(1.1)(*0;(4,3)*),^,(*0;(-2,-1)*){-}}
4790 \POS0\ar@{->}(-3,5)
4791 \POS0\ar@{->}(-3,4)
4792 \POS0,{\ellipse(1)(*0;(-5,-3)*),^,(*0;(-4,-3)*){-}}
4793 \end{xy}
4794 \caption{The cone of directions $C_{4,1}$}
4795 \label{f:C:4:1}
4796 \end{figure}
4798 Let us now consider the directions in which
4799 $\vec v_2$ is minimal while $\vec v_1$ is maximal.
4800 We find
4802 C_{2,1} = \poshull \,\{ (4,-5), (3,-4) \} \setminus \{ \vec 0 \}
4805 as shown in \autoref{f:C:2:1}.
4806 The vertices of the integer hull of $C_{2,1}$ are $(4,-5)$
4807 and $(3,-4)$.
4808 The corresponding widths are
4810 \begin{aligned}
4811 \vec c_1 &= (4,-5) & w_{\vec c_1} &= 6 \\
4812 \vec c_2 &= (3,-4) & w_{\vec c_2} &= 4
4814 \end{aligned}
4816 We similarly find
4818 C_{3,1} = \poshull \,\{ (4,-5), (-1,2) \} \setminus \{ \vec 0 \}
4821 with integer hull
4822 $\poshull \,\{ (4,-5), (-1,2), (1,-1) \}$, shown
4823 in \autoref{f:C:3:1}, yielding
4825 \begin{aligned}
4826 \vec c_3 &= (4,-5) & w_{\vec c_3} &= 6 \\
4827 \vec c_4 &= (-1,2) & w_{\vec c_4} &= 3 \\
4828 \vec c_5 &= (1,-1) & w_{\vec c_5} &= 3
4830 \end{aligned}
4832 On the other hand,
4834 C_{4,1} = \emptyset
4837 as shown in \autoref{f:C:4:1} and so this combination
4838 does not yield any width direction candidates.
4839 The other pairs of vertices further yield
4841 \begin{aligned}
4842 \vec c_6 &= (-1,2) & w_{\vec c_6} &= 3 \\
4843 \vec c_7 &= (-3,5) & w_{\vec c_7} &= 5 \\
4844 \vec c_8 &= (-3,4) & w_{\vec c_8} &= 4 \\
4845 \vec c_9 &= (-3,5) & w_{\vec c_9} &= 5 \\
4846 \vec c_{10} &= (-2,3) & w_{\vec c_{10}} &= 3
4848 \end{aligned}
4850 Since the polytope under consideration is not parametric,
4851 there is only one (non-empty, $0$-dimensional) chamber and
4852 it corresponds to one of the directions, say $\vec c_4 = (-1,2)$,
4853 with width $3$ (the other directions with the same width
4854 having been removed).
4856 \begin{figure}
4857 \intercol=1.1cm
4858 \begin{xy}
4859 <\intercol,0pt>:<0pt,\intercol>::
4860 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
4861 \POS0,{\xylattice{-0}{10}00}%
4862 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(10.5,0)}%
4863 \POS0,{\xylattice00{-0}7}%
4864 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*[|(2)]\xypolyline{}}
4865 \POS(-0.5,-0.5)\ar@{.}(7.5,7.5)
4866 \POS(0.5,-0.5)\ar@{.}(8.5,7.5)
4867 \POS(1.5,-0.5)\ar@{.}(9.5,7.5)
4868 \POS(2.5,-0.5)\ar@{.}(10.5,7.5)
4869 \POS(-0.5,-0.25)\ar@{-}(10.5,5.25)
4870 \POS(-0.5,0.25)\ar@{-}(10.5,5.75)
4871 \POS(-0.5,0.75)\ar@{-}(10.5,6.25)
4872 \POS(-0.5,1.25)\ar@{-}(10.5,6.75)
4873 \POS(-0.25,-0.5)\ar@{--}(10.5,6.666)
4874 \POS(-0.5,-0.333)\ar@{--}(10.5,7)
4875 \POS(-0.5,0)\ar@{--}(10.5,7.333)
4876 \POS(-0.5,0.333)\ar@{--}(10.25,7.5)
4877 \POS(0,0)*{\bullet}
4878 \POS(5,3)*{\bullet}
4879 \POS(5,4)*{\bullet}
4880 \POS(9,6)*{\bullet}
4881 \POS(3,2)*{\bullet}
4882 \POS(4,3)*{\bullet}
4883 \POS(6,4)*{\bullet}
4884 \POS(7,5)*{\bullet}
4885 \end{xy}
4886 \caption{A polytope and its lattice width directions}
4887 \label{f:width:2}
4888 \end{figure}
4890 Each of the three directions that yield the minimal
4891 width of 3 is shown in \autoref{f:width:2}.
4892 \end{example}
4894 \begin{example} \label{ex:width:2}
4895 Consider the polytope
4897 P(p) = \left\{\,
4898 \vec x \mid
4899 \begin{aligned}
4900 -2 x_1 + p + 5 &\ge 0 \\
4901 2 x_1 + p + 5 &\ge 0 \\
4902 -2 x_2 - p + 5 &\ge 0 \\
4903 2 x_2 - p + 5 &\ge 0
4904 \end{aligned}
4905 \,\right\}
4907 from \shortciteN[Example~2.1.7]{Woods2004PhD}.
4908 The parametric vertices are
4910 \begin{aligned}
4911 \vec v_1(p) & = \left(\frac{p+5}2, \frac{-p+5}2\right) \\
4912 \vec v_2(p) & = \left(\frac{p+5}2, \frac{p-5}2\right) \\
4913 \vec v_3(p) & = \left(\frac{-p-5}2, \frac{-p+5}2\right) \\
4914 \vec v_4(p) & = \left(\frac{-p-5}2, \frac{p-5}2\right)
4916 \end{aligned}
4918 We find two essentially different candidate width directions
4920 \begin{aligned}
4921 \vec c_1 &= (0,1) & w_{\vec c_1}(p) &= 5-p \\
4922 \vec c_2 &= (1,0) & w_{\vec c_2}(p) &= 5+p
4924 \end{aligned}
4926 The first direction can be found by combining, say,
4927 $\vec v_1(p)$ and $\vec v_2(p)$, while the second direction can be
4928 found by combining, say, $\vec v_1(p)$ and $\vec v_3(p)$.
4929 The parameter domain for the parametric polytope $P(p)$ is
4931 Q = \{\, p \mid -5 \le p \le 5 \,\}
4934 The two (closed) chambers are therefore
4936 \begin{aligned}
4937 C_1 &= \{\, p \in Q \mid 5 - p \le 5+p \,\} \\
4938 C_2 &= \{\, p \in Q \mid 5 + p \le 5-p \,\}
4940 \end{aligned}
4942 To obtain a partition, \autoref{s:interior} gives
4943 the internal point $(0,0)$, which happens to meet
4944 the facets $p \ge 0$ and $-p \ge 0$. We therefore
4945 keep the facet with positive (inner) normal closed
4946 and open up the other facet. The result is
4948 \begin{aligned}
4949 \hat C_1 &= \{\, p \mid 0 \le p \le 5 \,\} \\
4950 \hat C_2 &= \{\, p \mid -5 \le p < 0 \,\}
4952 \end{aligned}
4954 Since we are usually only interested in integer parameter
4955 values, the latter chamber would become
4956 $\hat C_2 = \{\, p \mid -5 \le p \le -1 \,\}$.
4957 \end{example}
4959 Our description differs slightly from that of
4960 of \shortciteN{Eisenbrand2007parameterised}.
4961 First, they consider all pairs of basic solutions instead
4962 of all pairs of vertices, which means that they may
4963 consider basic solutions that are never feasible and that,
4964 in case of a non-simple polytope,
4965 they may consider the same parametric vertex more than once.
4966 The set of integer
4967 directions for a pair of vertices is the intersection of
4968 the sets of integer directions they obtain for each of
4969 the corresponding basic solutions.
4970 Second, they use a different method of creating a partition
4971 of partially-open chambers, which may lead to some lower-dimensional
4972 chambers surviving and hence to a larger total number of chambers.
4975 \subsection{Testing whether a set has an infinite number of points}
4976 \label{s:infinite}
4978 In some situation we are given the generating function of
4979 some integer set and we would like to know if the set is
4980 infinite or not. Typically, we want to know if the set
4981 is empty or not, but we cannot simply count the number of elements
4982 in the standard way since we may not have any guarantee that
4983 the set has only a finite number of elements.
4984 We will consider the slightly more general case where we are
4985 given a rational generating function $f(\vec x)$ of the form~\eqref{eq:rgf}
4986 such that
4987 \begin{equation}
4988 \label{eq:rgf:psp}
4989 f(\vec x) = \sum_{\vec s \in Q \cap \ZZ^d} c(\vec s)\, \vec x^{\vec s}
4990 \end{equation}
4991 converges on some nonempty open subset of $\CC^d$, $Q$ is a pointed
4992 polyhedron and $c(\vec s) \ge 0$,
4993 and we want to compute
4994 \begin{equation}
4995 \label{eq:psp:sum}
4996 S = \sum_{\vec s \in Q \cap \ZZ^d} c(\vec s)
4998 \end{equation}
4999 where the sum may diverge, i.e., ``$S = \infty$''.
5000 The following proposition shows that we can determine $S$
5001 in polynomial time.
5002 For a sketch of an alternative technique, see
5003 \shortciteN[Proof of Lemma~16]{Woods2005period}.
5005 \begin{proposition}
5006 Fix $d$ and $k$.
5007 Given a \rgf/ of the form~\eqref{eq:rgf} with $k_i \le k$
5008 and a pointed polyhedron $Q \subset \QQ^d$, then there is a
5009 polynomial time algorithm that determines for the corresponding
5010 function $c(\vec s)$~\eqref{eq:rgf:psp} whether the sum~\eqref{eq:psp:sum}
5011 diverges and computes the value of $S$~\eqref{eq:psp:sum} if it does not.
5012 \end{proposition}
5013 \begin{proof}
5014 Since $Q$ is pointed, the series~\eqref{eq:rgf:psp} converges on a neighborhood
5015 of $e^{\vec \ell} = (e^{\ell_1}, \ldots, e^{\ell_d})$ for any $\vec \ell$
5016 such that $\sps {\vec r_k} {\vec \ell} < 0$ for
5017 any (extremal) ray $\vec r_k$ of $Q$ and
5018 such that $\sps {\vec b_{i j}} {\vec \ell} \ne 0$ for any
5019 $\vec b_{i j}$ in~\eqref{eq:rgf}.
5020 Let $\vec \alpha = - \vec \ell$ and perform the substitution
5021 $\vec x = t^{\vec \alpha}$. The function $g(t) = f(t^{\vec \alpha})$
5022 is then also a (short) \rgf/ and
5024 g(t) = \sum_{k \in \sps {\vec\alpha} Q \cap \ZZ}
5025 \left(
5026 \sum_{\shortstack{$\scriptstyle \vec s \in Q \cap \ZZ^d$\\
5027 $\scriptstyle \sp \alpha s = k$}} c(\vec s)
5028 \right) t^k
5029 =: \sum_{k \in \sps {\vec\alpha} Q \cap \ZZ} d(k) \, t^k
5032 with $\sps {\vec\alpha} Q = \{ \sp \alpha x \mid \vec x \in Q \}$,
5033 converges in a neighborhood of $e^{-1}$, while
5035 S = \sum_{k \in \sps {\vec\alpha} Q \cap \ZZ} d(k)
5038 Since $c(\vec s) \ge 0$, we have $d(k) \ge 0$
5039 and the above sum diverges iff any of the coefficients of the
5040 negative powers of $t$ in the Laurent expansion of $g(t)$ is non-zero.
5041 If the sum converges, then the sum is simply the coefficient
5042 of the constant term in this expansion.
5044 It only remains to show now that we can compute a suitable $\vec \alpha$
5045 in polynomial time, i.e., an $\vec \alpha$ such that
5046 $\sps {\vec r_k} {\vec \alpha} > 0$ for any (extremal) ray $\vec r_k$ of $Q$ and
5047 $\sps {\vec b_{i j}} {\vec \alpha} \ne 0$ for any
5048 $\vec b_{i j}$ in~\eqref{eq:rgf}.
5049 By adding the $\vec r_k$ to the list of $\vec b_{i j}$ if needed, we can relax
5050 the first set of constraints to $\sps {\vec r_k} {\vec \alpha} \ge 0$.
5051 Let $Q$ be described by the constraints $A \vec x + \vec c \ge \vec 0$
5052 and let $B$ be $d \times d$ non-singular submatrix of $A$, obtained
5053 by removing some of the rows of $A$. Such a $B$ exists since
5054 $Q$ does not contain any straight line.
5055 Clearly, $B \vec r \ge \vec 0$ for any ray $\vec r$ of $Q$.
5056 Let $\vec b'_{i j} = B \vec b_{i j}$, then since $\vec b_{i j} \ne \vec 0$
5057 and B is non-singular, we have $\vec b'_{i j} \ne \vec 0$.
5058 We may therefore find in polynomial time a point $\vec \alpha' \ge \vec 0$
5059 on the ``\ai{moment curve}'' such that
5060 $\sps {\vec b'_{i j}} {\vec \alpha'} \ne 0$
5061 \shortcite[Algorithm~5.2]{Barvinok1999}.
5062 Let $\vec \alpha = B^\T \vec \alpha'$.
5063 Then
5065 \sps {\vec b_{i j}} {\vec \alpha}
5067 \sps {\vec b_{i j}} {B^\T \vec \alpha'}
5069 \sps {B \vec b_{i j}} {\vec \alpha'}
5071 \sps {\vec b'_{i j}} {\vec \alpha'}
5072 \ne 0
5076 \sps {\vec r_k} {\vec \alpha}
5078 \sps {\vec r_k} {B^\T \vec \alpha'}
5080 \sps {B \vec r_k} {\vec \alpha'}
5081 \ge 0
5084 as required.
5085 Note that in practice, we would, as usual, first try a
5086 fixed number of random vectors $\vec \alpha' \ge \vec 0$
5087 before resorting to looking for a point on the moment curve.
5088 \end{proof}
5091 \subsection{Enumerating integer projections of parametric polytopes}
5092 \label{s:projection}
5094 In this section we are interested in computing
5095 \begin{equation}
5096 \label{eq:count:projection}
5097 c(\vec s)=\#\left\{\vec t\in\ZZ^{d} \mid \exists \vec u\in\ZZ^{m}:
5098 (\vec s,\vec t,\vec u)\in P\right\}
5100 \end{equation}
5101 with $P \subset \QQ^{n}\times\QQ^{d}\times\QQ^{m}$ a rational
5102 pointed polyhedron such that
5104 P_{\vec s}=\left\{(\vec t,\vec u)\in\QQ^d\times\QQ^m
5105 \mid (\vec s,\vec t,\vec u)\in P\right\}
5107 is a polytope for any $\vec s$.
5108 This is equivalent to computing the number of points
5109 in the \ai{integer projection} of a parametric polytope
5111 c(\vec s)=\#\big(\pi(P_{\vec s}\cap\ZZ^{d+m})\big)
5114 with $\pi:\QQ^d\times\QQ^m\rightarrow\QQ^d$ defined by
5115 $\pi(\vec t, \vec u)=\vec t$.
5116 Exponential methods for computing $c(\vec s)$ are
5117 described by \shortciteN{Verdoolaege2005experiences}
5118 and \shortciteN{Seghir2006memory}.
5119 Here, we provide some implementation details for the polynomial
5120 method of \shortciteN[Theorem~1.7]{Woods2003short}, for
5121 computing the generating function, $\sum_{\vec s}c(\vec s) \, \vec x^{\vec s}$,
5122 which can then be converted into an explicit function $c(\vec s)$
5123 \shortcite[Corollary~1.11]{Verdoolaege2008counting}.
5124 Note that in contrast to \shortciteN[Theorem~1.7]{Woods2003short},
5125 we may allow $P$ to be an unbounded (but still pointed) polyhedron here
5126 (as long as $P_{\vec s}$ is bounded), since
5127 we replace their application of
5128 \shortciteN[Lemma~3.1]{Kannan1992}
5129 by \shortciteN[Theorem~5]{Eisenbrand2007parameterised}.
5131 If there is only one existentially quantified variable ($m = 1$),
5132 then computing~\eqref{eq:count:projection} is easy.
5133 You simply shift $P$ by $1$ in the $u$ direction and subtract
5134 this shifted copy from the original,
5136 D = P \setminus (\vec e_{n+d+1} + P)
5139 (See, e.g., \shortciteN[Figure~1, page~973]{Woods2003short}
5140 or \shortciteN[Figure~4.33, page~186]{Verdoolaege2005PhD}.)
5141 In the difference $D$ there will be {\em exactly} one value of $u$
5142 for each value of the remaining variables for which there was
5143 {\em at least} one value of $u$ in $P$,
5145 \forall (\vec s, \vec t):
5146 \quad
5147 \left(
5148 \exists u: (\vec s, \vec t, u) \in P
5149 \right)
5150 \iff
5151 \left(
5152 \exists! u: (\vec s, \vec t, u) \in D
5153 \right)
5156 The function $c(\vec s)$ can then be computed by counting
5157 the number of elements in $D(\vec s)$.
5158 These operations can be performed either in the space
5159 of (unions of) parametric polytopes or on generating functions.
5160 In the first case, $D(\vec s)$ can be written as a disjoint union
5161 of parametric polytopes that can be enumerated separately.
5162 In the second case, we first compute the generating function
5163 $f(\vec x, \vec y)$ of the set
5167 (\vec s, \vec t) \mid \exists u \in \ZZ : (\vec s, \vec t, u) \in P
5170 and then obtain the generating function $C(\vec x)$ of $c(\vec s)$
5171 as $C(\vec x) = f(\vec x, \vec 1)$.
5172 In the remainder of this section, we will concentrate on the
5173 computation of the generating function of $S$.
5174 To compute this generating function in the current case where
5175 there is only one existentially quantified variable, we first
5176 compute the generating function $g(\vec x, \vec y, z)$ of
5177 $P(\vec s, \vec t, u)$, perform operations on the generating function
5178 equivalent to the set operations (see, e.g.,
5179 \shortciteN[Section~4.5.3]{Verdoolaege2005PhD}), resulting
5180 in a generating function $g'(\vec x, \vec y, z)$, and then sum
5181 over all values (at most one for each value of $\vec s$
5182 and $\vec t$) of $u$, i.e., $f(\vec x, \vec y) = g'(\vec c, \vec y, 1)$.
5184 \begin{figure}
5185 \intercol=1.05cm
5186 \begin{xy}
5187 <\intercol,0pt>:<0pt,\intercol>::
5188 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
5189 \POS0,{\xylattice{-0}{10}00}%
5190 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(10.5,0)}%
5191 \POS0,{\xylattice00{-0}7}%
5192 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*\xypolyline{}}
5193 \POS(0,0)*[*0.333]\xybox{
5194 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*\xypolyline{--}}
5196 \POS(0,0)*{\bullet}
5197 \POS(5,3)*{\bullet}
5198 \POS(5,4)*{\bullet}
5199 \POS(9,6)*{\bullet}
5200 \POS(3,2)*{\bullet}
5201 \POS(4,3)*{\bullet}
5202 \POS(6,4)*{\bullet}
5203 \POS(7,5)*{\bullet}
5204 \POS(-1,-0.5)\ar@{-}(-1,7.5)
5205 \POS(-1,0)*{\bullet}
5206 \POS(-1,3)*{\bullet}
5207 \POS(-1,4)*{\bullet}
5208 \POS(-1,6)*{\bullet}
5209 \POS(-1,2)*{\bullet}
5210 \POS(-1,3)*{\bullet}
5211 \POS(-1,4)*{\bullet}
5212 \POS(-1,5)*{\bullet}
5213 \POS(-0.5,-1)\ar@{-}(10.5,-1)
5214 \POS(0,-1)*+++!UR{S}
5215 \POS(0,-1)*{\bullet}
5216 \POS(5,-1)*{\bullet}
5217 \POS(5,-1)*{\bullet}
5218 \POS(9,-1)*{\bullet}
5219 \POS(3,-1)*{\bullet}
5220 \POS(4,-1)*{\bullet}
5221 \POS(6,-1)*{\bullet}
5222 \POS(7,-1)*{\bullet}
5223 \end{xy}
5224 \caption{A polytope and its integer projections}
5225 \label{f:projection}
5226 \end{figure}
5228 If there is more than one existentially quantified variable ($m > 1$),
5229 then we can in principle apply the above shifting and subtracting
5230 technique recursively to obtain a generating function
5231 $f(\vec x, \vec y)$ for the set
5232 \begin{equation}
5233 \label{eq:projection:T}
5236 (\vec s, \vec t) \mid \exists \vec u \in \ZZ^m : (\vec s, \vec t, \vec u) \in P
5238 \end{equation}
5239 and then compute $C(\vec x) = f(\vec x, \vec 1)$.
5240 There are however some complications.
5241 Most notably, after applying the technique in one direction
5242 and projecting out the corresponding variable, the resulting set, i.e.,
5246 (\vec s, \vec t, u_1, \ldots, u_{m-1}) \mid
5247 \exists u_m \in \ZZ : (\vec s, \vec t, \vec u) \in P
5251 in general does not correspond to the integer points in some polytope.
5252 For example, assume that the polytope in \autoref{f:projection}
5253 contains the values of $\vec u$ associated to a particular
5254 value of $(\vec s, \vec t)$. Since there are integer points
5255 in this polytope, we should count this value of $\vec t$, but only once.
5256 If we apply the above technique in the vertical direction ($u_2$), then
5257 we can compute (a generating function for) the set $S$ shown
5258 on the bottom of the figure.
5259 Note, however, that there are ``gaps'' in this set, i.e.,
5260 if we compute $S \setminus (\vec e_{n+d+1} + S)$ then we will not
5261 end up with a single point (for this value of $(\vec s, \vec t)$).
5262 Since the biggest gap is three wide, we need
5263 to compute
5266 \setminus (\vec e_{n+d+1} + S)
5267 \setminus (2 \vec e_{n+d+1} + S)
5268 \setminus (3 \vec e_{n+d+1} + S)
5270 to obtain a single point.
5271 If we do the subtraction in the horizontal direction first,
5272 then we end up with a set (shown on the left) with gaps
5273 at most two wide, so afterwards we only need to subtract twice in the
5274 vertical direction.
5276 \begin{figure}
5277 \intercol=0.8cm
5278 \begin{xy}
5279 <\intercol,0pt>:<0pt,\intercol>::
5280 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
5281 \POS0,{\xylattice{-0}{4}00}%
5282 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(4.5,0)}%
5283 \POS0,{\xylattice00{-0}7}%
5284 \POS@i@={(0,0),(1,3),(3,6),(3,4),(0,0)},{0*\xypolyline{}}
5285 \POS(0,0)*{\bullet}
5286 \POS(1,3)*{\bullet}
5287 \POS(3,4)*{\bullet}
5288 \POS(3,6)*{\bullet}
5289 \POS(1,2)*{\bullet}
5290 \POS(2,3)*{\bullet}
5291 \POS(2,4)*{\bullet}
5292 \POS(3,5)*{\bullet}
5293 \POS(-0.5,-1)\ar@{-}(4.5,-1)
5294 \POS(0,-1)*+++!UR{S}
5295 \POS(0,-1)*{\bullet}
5296 \POS(1,-1)*{\bullet}
5297 \POS(2,-1)*{\bullet}
5298 \POS(3,-1)*{\bullet}
5299 \end{xy}
5300 \caption{A transformed polytope and its integer projection}
5301 \label{f:projection:2}
5302 \end{figure}
5304 In general, there is no bound on the widths of the gaps we may
5305 encounter in any given direction. However, there are directions
5306 in which the gaps are known to be ``small''.
5307 In particular, if the dimension $m$ is fixed, then the lattice width
5308 (see \autoref{s:width}) of lattice point free polytopes
5309 is bounded by a constant $\omega(m)$%
5310 ~\shortcite{Lagarias90,Barvinok02,Banaszczyk1999flatness}.
5311 This means that in the direction of the lattice width of a polytope,
5312 the gaps will be not be larger than $\omega(m)$
5313 \shortcite[Theorem~4.3]{Woods2003short}.
5314 Otherwise, we would be able to put a (uniformly) scaled down version
5315 of the polytope in the gap and it would contain no lattice points,
5316 which would contradict the fact that its lattice width is bounded
5317 by $\omega(m)$.
5318 \autoref{f:projection} contains such a scaled down copy
5319 of the original polytope. However, neither the horizontal
5320 nor the vertical direction is a lattice width direction
5321 of this polytope. The actual lattice width of this
5322 polytope was computed in Example~\ref{ex:width} as $3$
5323 with corresponding direction $\vec c = (-1,2)$.
5324 \autoref{f:projection:2} shows the result of applying
5325 the unimodular transformation
5327 \begin{bmatrix}
5328 -1 & 2 \\
5329 0 & 1
5330 \end{bmatrix}
5332 to the polytope. Note that the horizontal direction
5333 now has gaps of width at most 1. After shifting, subtracting
5334 and projecting in the vertical direction, we therefore
5335 end up with a set $S$ with gaps of width 1 and we then
5336 only have to shift and subtract once in the remaining
5337 (horizontal) direction.
5339 In fact, for two-dimensional polytopes the gaps in the lattice
5340 width direction will always be one, as shown by the following lemma.
5341 \begin{lemma}
5342 \label{l:gap}
5343 For any rational polygon, the gaps in a lattice
5344 width direction are of width at most 1.
5345 \end{lemma}
5346 \begin{proof}
5347 We may assume that $x$ is the given lattice width direction of
5348 a given polygon $P$.
5349 If there is a gap of width 2, then there is an integer value $x_1$ of $x$
5350 such that
5351 $P \cap \{\, (x_1, y) \,\} \ne \emptyset$,
5352 $P \cap \{\, (x_1+2, y) \,\} \ne \emptyset$,
5353 while
5354 $P \cap \{\, (x_1+1, y) \,\} \cap \ZZ^2 = \emptyset$.
5355 Using \shortciteN[Lemma~4.2]{Woods2003short}, we can put
5356 a scaled down copy $P'$ of $P$ between $x=x_1$ and $x=x_1+2$
5357 (and inside of $P$).
5358 $P'$ meets the line $x=x_1+1$ between two consecutive integer
5359 points, $y_1$ and $y_1+1$. Let $P''$ be the polygon bounded by $x=x_1$ and
5360 $x=x_1+2$ and two lines that separate $P'$ from these two
5361 integer points.
5362 $P''$ will have the same width (2) in the
5363 $x$ direction, while $P' \subset P''$.
5364 The $x$ direction is therefore also a lattice width direction of $P''$.
5365 $P''$ cannot intersect both $x=x_1$ and $x=x_1+2$ in a segment of
5366 length greater than or equal to $1$.
5367 Otherwise, it would also intersect $x=x_1+1$ in a segment of length
5368 greater than or equal to $1$.
5370 We may therefore assume that the length of the intersection
5371 of $P''$ with $x=x_1$ is smaller than $1$.
5372 If this line segment contains an integer point, then call it $y_2$.
5373 Otherwise, let $y_2$ be the greatest integer smaller than the
5374 points in the line segment.
5375 We may assume that $y_1 = y_2$.
5376 Otherwise, we can apply the unimodular transformation
5378 \begin{bmatrix}
5379 x \\
5381 \end{bmatrix}
5383 \begin{bmatrix}
5384 1 & 0 \\
5385 y_1 - y_2 & 1
5386 \end{bmatrix}
5387 \begin{bmatrix}
5388 x \\
5390 \end{bmatrix}
5393 without changing the width in direction $x$.
5394 If $P''$ contains $(x_1, y_1)$, it intersects $x=x_1$
5395 in a segment $[y_1-\alpha_1, y_1+\alpha_2]$.
5396 We may then similarly assume that $\alpha_2 \ge \alpha_1$.
5397 $P''$ will only cut $x=x_1+2$ in points with $y$-coordinate
5398 smaller than $2-\alpha_2$. The width in the $y$ direction
5399 will therefore be smaller than $2-\alpha_2+\alpha_1 \le 2$,
5400 contradicting that $x$ is a lattice width direction.
5401 If $P''$ does not contain $(x_1, y_1)$, then it only
5402 intersects $x=x_1$ in points with $y$-coordinate $y_1+\alpha$
5403 with $0 < \alpha < 1$. Given any such point, it is clear
5404 that $P''$ intersects $x=x_1+2$ only in points with $y$-coordinate
5405 strictly between $y_1-\alpha$ and $y_1+1-\alpha$, again
5406 showing that the width in the $y$ direction is smaller than $2$ and
5407 leading to the same contradiction.
5408 The contradiction shows that there can be no gaps
5409 of width 2 in the lattice width direction of $P$.
5410 \end{proof}
5411 Note that the $\omega(2)$ bound is too coarse to reach
5412 the above conclusion as $\omega(2) > 2$.
5413 An example of a polygon with lattice with greater than $2$ is the polygon
5414 with vertices $(-17/110,83/110)$, $(2/10,-9/10)$ and $(177/90, 100/90)$,
5415 shown in \autoref{f:empty:width:2}, which has width $221/110$.
5417 \begin{figure}
5418 \intercol=3cm
5419 \begin{xy}
5420 <\intercol,0pt>:<0pt,\intercol>::
5421 \def\latticebody{\POS="c"+(0,-1.5)\ar@{--}"c"+(0,1.5)}%
5422 \POS0,{\xylattice{-1}{2}00}%
5423 \def\latticebody{\POS="c"+(-1.5,0)\ar@{--}"c"+(2.5,0)}%
5424 \POS0,{\xylattice00{-1}1}%
5425 \POS@i@={(-0.1545,0.7545),(0.2,-0.9),(1.966,1.111),(-0.1545,0.74545)},{0*\xypolyline{}}
5426 \end{xy}
5427 \caption{Lattice point free polygon with lattice width 2}
5428 \label{f:empty:width:2}
5429 \end{figure}
5431 The idea of the projection algorithm
5432 is now to first find a direction in which the gaps
5433 are expected to be small and to unimodularly transform
5434 the existentially quantified variables such that this direction
5435 lies in the direction of one of the transformed variables.
5436 Then, the remaining existentially quantified variables are
5437 projected out by applying the technique recursively.
5438 The resulting generating function will have gaps at most
5439 $\omega(m)$ wide and so we have to subtract at most
5440 $\omega(m)$ shifted copies of this generating function
5441 before we can plug in 1 to project out the selected
5442 (and now only remaining) existentially quantified variable.
5443 We now look at each of these step in a bit more detail.
5445 We are given a polyhedron $P$ such that $P_{\vec s}$ is a polytope
5446 and we want to compute a generating function $f(\vec x, \vec y)$
5447 for the set $T$ in~\eqref{eq:projection:T}.
5448 We first compute the lattice width directions of
5449 the $m$-dimensional parametric polytope $P_{\vec s, \vec t}$
5450 as in \autoref{s:width}.
5451 The result is a partition of the parameter domain, i.e.,
5452 the projection of $P$ onto the first $n+d$ coordinates,
5453 into partially open polyhedra $Q_i$, together with
5454 the lattice width direction $\vec c_i$ corresponding to each $Q_i$.
5455 Since the generating functions only encode integer points,
5456 we can replace each open facet $\sp a x + b > 0$ by the closed
5457 facet $\sp a x + b - 1 \ge 0$ to obtain a collection of closed
5458 polyhedra $\tilde Q_i$. Now let
5460 P_i = P \cap \tilde Q_i \times \QQ^m
5462 and let $f_i(\vec x, \vec y)$ be the generating function of the set
5464 T_i =
5466 (\vec s, \vec t) \mid
5467 \exists \vec u \in \ZZ^m : (\vec s, \vec t, \vec u) \in P_i
5471 Then clearly,
5473 f(\vec x, \vec y) = \sum_i f_i(\vec x, \vec y)
5476 From now on, we will consider a particular $P_i$ with corresponding
5477 lattice width $\vec c_i$ and drop the $i$ subscript.
5479 We are now given a polyhedron $P$ such that the lattice width
5480 direction of $P_{\vec s, \vec t}$ is $\vec c$.
5481 We first extend $\vec c$ to an $m \times m$ unimodular matrix $U$
5482 using the technique of \autoref{s:completion},
5486 \begin{bmatrix}
5487 \vec c^\T
5490 \end{bmatrix}
5492 and then compute
5494 P' =
5495 \begin{bmatrix}
5496 I_n & 0 & 0 \\
5497 0 & I_d & 0 \\
5498 0 & 0 & U
5499 \end{bmatrix}
5503 We have
5507 (\vec s, \vec t) \mid
5508 \exists \vec u' \in \ZZ^m : (\vec s, \vec t, \vec u') \in P'
5512 i.e., we may have changed the values of the existentially
5513 quantified variables, but we have not changed the set $T$.
5514 Now consider the set
5516 T' =
5518 (\vec s, \vec t, u_1') \mid
5519 \exists (u_2',\ldots,u_m') \in \ZZ^{m-1} :
5520 (\vec s, \vec t, \vec u') \in P'
5524 This set has only $m-1$ existentially quantified variables, so
5525 we may apply this projection algorithm recursively and obtain
5526 the generating function $f'(\vec x, \vec y, z)$ for $T'$.
5527 The set $T'$ may no longer correspond to the integer points
5528 in a polytope, but, by construction, the gaps in the final
5529 coordinate are small ($\le \omega(m)$).
5531 By now we have a generating function $f'(\vec x, \vec y, z)$
5532 for the set $T'$ (with small
5533 gaps in the final coordinate) and we have to compute the
5534 generating function $f(\vec x, \vec y)$ for $T$.
5535 By computing
5536 \begin{equation}
5537 \label{eq:projection:omega}
5538 f''(\vec x, \vec y, z) =
5539 f'(\vec x, \vec y, z) \bigoplus_{k=1}^{\floor{\omega(m)}}
5540 \left( z^k f'(\vec x, \vec y, z) \right)
5542 \end{equation}
5543 where $\oplus$ represents the operation on generating functions
5544 that corresponds to set difference on the corresponding sets,
5545 we obtain a generating for the set $T''$ where only
5546 the smallest value of $u_1'$ is retained.
5547 The total number of $u_1'$s associated to any $(\vec s, \vec t)$
5548 is therefore either zero or one and so the ``multiset'' defined
5549 by taking as many copies of $(\vec s, \vec t)$ as there are
5550 associated values of $u_1'$ is actually the set $T$.
5551 That is
5553 f(\vec x, \vec y) = f''(\vec x, \vec y, 1)
5557 The only remaining problem is that the ``$\oplus$'' operation
5558 in~\eqref{eq:projection:omega} is fairly expensive.
5559 In particular, this operation is performed by first
5560 computing the \ai{Hadamard product} of the two generating functions
5561 (which corresponds to the intersection of the sets) and
5562 then subtracting the resulting generating function from this
5563 first generating function.
5564 The last operation is fairly cheap, but the Hadamard product
5565 has a time complexity which while polynomial if the dimension (in
5566 this case the maximum of $k_i$ in~\eqref{eq:rgf}) is fixed,
5567 is exponential in this dimension.
5568 Furthermore, this dimension increases by $\max k_i - d$ on each
5569 successive application of the Hadamard product, while $\max k_i > d$
5570 as soon as some projection is performed, which certainly happens in the
5571 recursive application of the algorithm.
5572 Now, the total number of Hadamard products is bounded by a constant
5573 $\floor{\omega(m)}$ and so the increase in dimension is also bounded
5574 by a constant, so the whole operation is still polynomial for
5575 fixed dimension.
5576 Nevertheless, we do not want to perform any additional Hadamard
5577 products if we do not really have to.
5578 That is, we would like to be able to detect when we can stop
5579 performing these operations {\em before} reaching the upper
5580 bound $\floor{\omega(m)}$.
5582 Let $f'_0(\vec x, \vec y, z) = f'(\vec x, \vec y, z)$ and
5583 let $f'_k(\vec x, \vec y, z)$ be the result of applying
5584 the ``set difference'' in~\eqref{eq:projection:omega} $k$ times.
5585 Denote the corresponding sets by $T'_0$ and $T'_k$.
5586 We want to find the smallest $k$ such that
5587 $f''(\vec x, \vec y, z) = f'_k(\vec x, \vec y, z)$.
5588 Note that we are talking about equality of rational functions here.
5589 Any further application of the set difference operation will
5590 not change this rational function, but it {\em will\/} typically
5591 produce a different (more complex) representation.
5592 To check whether the current $k$ is sufficient, we are going to
5593 count how many times any element of $T'_k$ still appears in a
5594 shifted copy of $T'_0$, with shift greater than or equal to $k+1$.
5595 If this number is zero, any further set difference will have no effect.
5596 That is, we want to compute
5598 \sum_{l=k+1}^\infty
5599 \left|
5600 T'_l \cap \left(\vec e_{n+d+1} + T' \right)
5601 \right|
5604 or, in terms of generating functions,
5606 h(\vec x, \vec y, z) = \sum_{l=k+1}^\infty
5607 f_k'(\vec x, \vec y, z) \star z^l \, f'(\vec x, \vec y, z)
5610 We should point out here that while the Hadamard product ($\star$)
5611 corresponds to intersection when applied to generator functions
5612 of indicator functions (i.e., with coefficients one or zero),
5613 in general it will produce a generating function with coefficients
5614 that are the product of the corresponding coefficients in the two
5615 operands.
5616 We can therefore rewrite the above equation as
5618 \begin{aligned}
5619 h(\vec x, \vec y, z) & = \sum_{l=k+1}^\infty
5620 f_k'(\vec x, \vec y, z) \star z^l \, f'(\vec x, \vec y, z)
5622 & = f_k'(\vec x, \vec y, z) \star
5623 \left(
5624 \sum_{l=k+1}^\infty z^l \, f'(\vec x, \vec y, z)
5625 \right)
5627 & = f_k'(\vec x, \vec y, z) \star
5628 \frac{z^{k+1} \, f'(\vec x, \vec y, z)}{1-z}
5630 \end{aligned}
5632 Computing $h(\vec x, \vec y, 1)$ would give us a generating
5633 function with as coefficients how many times a point of $T'_k$
5634 still appears in a shifted copy of $T'_0$ for any particular
5635 value of $\vec s$ and $\vec t$.
5636 However, we want to know if this number is zero for {\em all\/}
5637 values of $\vec s$ and $\vec t$, so we compute $h(\vec 1, \vec 1, 1)$
5638 instead. We have to be careful here since we allow the polyhedron
5639 $P$ to be unbounded and so we should apply the technique
5640 of \autoref{s:infinite} with $Q$ the projection of $P$ on
5641 the remaining coordinates.
5643 Note that testing whether we can stop is more expensive
5644 than applying the next iteration (since we have an extra
5645 $(1-z)$ factor in the denominator of one of the operands).
5646 However, we may save many iterations by stopping early
5647 and we will not needlessly replace a given representation
5648 of $f''(\vec x, \vec y, z)$ by a more complex representation
5649 (with more factors in the denominator).
5650 An alternative way of checking whether we can stop is to
5651 test whether $f'_k(\vec x, \vec y, z) = f'_{k+1}(\vec x, \vec y, z)$
5652 (as rational functions).
5653 To do so, we would need to check that both
5655 f'_k(\vec x, \vec y, z) -
5656 \left( f'_k(\vec x, \vec y, z) \star f'_{k+1}(\vec x, \vec y, z) \right)
5660 f'_{k+1}(\vec x, \vec y, z) -
5661 \left( f'_k(\vec x, \vec y, z) \star f'_{k+1}(\vec x, \vec y, z) \right)
5663 are zero and this Hadamard product is more expensive than
5664 the one above.
5666 \begin{figure}
5667 \intercol=1.05cm
5668 \begin{xy}
5669 <\intercol,0pt>:<0pt,\intercol>::
5670 \def\latticebody{\POS="c"+(0,-5.5)\ar@{--}"c"+(0,5.5)}%
5671 \POS0,{\xylattice{-5}{5}00}%
5672 \def\latticebody{\POS="c"+(-5.5,0)\ar@{--}"c"+(5.5,0)}%
5673 \POS0,{\xylattice00{-5}5}%
5674 \POS(0,-5.5)\ar(0,5.5) \POS(0,5.5)*+!UL{x_2}
5675 \POS(-5.5,0)\ar(5.5,0) \POS(5.5,0)*+!DR{x_1}
5676 \POS@i@={(-5,0),(5,0)},{0*[|(2)]\xypolyline{}}
5677 \POS@i@={(-4.5,0.5),(4.5,0.5),(4.5,-0.5),(-4.5,-0.5),(-4.5,0.5)},{0*[|(2)]\xypolyline{}}
5678 \POS@i@={(-4,1),(4,1),(4,-1),(-4,-1),(-4,1)},{0*[|(2)]\xypolyline{}}
5679 \POS@i@={(-3.5,1.5),(3.5,1.5),(3.5,-1.5),(-3.5,-1.5),(-3.5,1.5)},{0*[|(2)]\xypolyline{}}
5680 \POS@i@={(-3,2),(3,2),(3,-2),(-3,-2),(-3,2)},{0*[|(2)]\xypolyline{}}
5681 \POS@i@={(-2.5,2.5),(2.5,2.5),(2.5,-2.5),(-2.5,-2.5),(-2.5,2.5)},{0*[|(2)]\xypolyline{}}
5682 \POS@i@={(-2,3),(2,3),(2,-3),(-2,-3),(-2,3)},{0*[|(2)]\xypolyline{}}
5683 \POS@i@={(-1.5,3.5),(1.5,3.5),(1.5,-3.5),(-1.5,-3.5),(-1.5,3.5)},{0*[|(2)]\xypolyline{}}
5684 \POS@i@={(-1,4),(1,4),(1,-4),(-1,-4),(-1,4)},{0*[|(2)]\xypolyline{}}
5685 \POS@i@={(-0.5,4.5),(0.5,4.5),(0.5,-4.5),(-0.5,-4.5),(-0.5,4.5)},{0*[|(2)]\xypolyline{}}
5686 \POS@i@={(0,-5),(0,5)},{0*[|(2)]\xypolyline{}}
5687 \POS(-5,0)*+!DR{5}
5688 \POS(-4.5,0.5)*+!DR{4}
5689 \POS(-4,1)*+!DR{3}
5690 \POS(-3.5,1.5)*+!DR{2}
5691 \POS(-3,2)*+!DR{1}
5692 \POS(-2.5,2.5)*+!DR{0}
5693 \POS(-2,3)*+!DR{-1}
5694 \POS(-1.5,3.5)*+!DR{-2}
5695 \POS(-1,4)*+!DR{-3}
5696 \POS(-0.5,4.5)*+!DR{-4}
5697 \POS(-0,5)*+!DR{-5}
5698 \end{xy}
5699 \caption{The parametric polytope from Example~\ref{ex:projection}
5700 for different values of the parameter}
5701 \label{f:ex:projection}
5702 \end{figure}
5704 \begin{example} \label{ex:projection}
5705 Consider once more the parametric polytope
5707 P(p) = \left\{\,
5708 \vec x \mid
5709 \begin{aligned}
5710 -2 x_1 + p + 5 &\ge 0 \\
5711 2 x_1 + p + 5 &\ge 0 \\
5712 -2 x_2 - p + 5 &\ge 0 \\
5713 2 x_2 - p + 5 &\ge 0
5714 \end{aligned}
5715 \,\right\}
5717 from \shortciteN[Example~2.1.7]{Woods2004PhD}
5718 and Example~\ref{ex:width:2} and assume we want to
5719 compute
5721 c(p) = \left[ \exists \vec x \in \ZZ^2: (p, \vec x) \in P \right]
5724 That is, we simply want to know for which values of $p$
5725 the polytope is non-empty.
5726 Now, there are more efficient ways of answering this particular question,
5727 but we will use it here as an example of the above algorithm.
5728 The polytope $P(p)$ is shown in \autoref{f:ex:projection} for
5729 all integer value of the parameter for which the polytope
5730 is non-empty.
5732 \begin{figure}
5733 \intercol=1.05cm
5734 \begin{xy}
5735 <\intercol,0pt>:<0pt,\intercol>::
5736 \def\latticebody{\POS="c"+(0,-5.5)\ar@{--}"c"+(0,5.5)}%
5737 \POS0,{\xylattice{-5}{5}00}%
5738 \def\latticebody{\POS="c"+(-5.5,0)\ar@{--}"c"+(5.5,0)}%
5739 \POS0,{\xylattice00{-5}5}%
5740 \POS(0,-5.5)\ar(0,5.5) \POS(0,5.5)*+!UL{x_2}
5741 \POS(-5.5,0)\ar(5.5,0) \POS(5.5,0)*+!DR{x_1}
5742 \POS@i@={(-2.5,2.5),(2.5,2.5),(2.5,-2.5),(-2.5,-2.5),(-2.5,2.5)},{0*[|(2)]\xypolyline{}}
5743 \POS@i@={(-2,3),(2,3),(2,-3),(-2,-3),(-2,3)},{0*[|(2)]\xypolyline{}}
5744 \POS@i@={(-1.5,3.5),(1.5,3.5),(1.5,-3.5),(-1.5,-3.5),(-1.5,3.5)},{0*[|(2)]\xypolyline{}}
5745 \POS@i@={(-1,4),(1,4),(1,-4),(-1,-4),(-1,4)},{0*[|(2)]\xypolyline{}}
5746 \POS@i@={(-0.5,4.5),(0.5,4.5),(0.5,-4.5),(-0.5,-4.5),(-0.5,4.5)},{0*[|(2)]\xypolyline{}}
5747 \POS@i@={(0,-5),(0,5)},{0*[|(2)]\xypolyline{}}
5748 \POS(-2.5,2.5)*+!DR{0}
5749 \POS(-2,3)*+!DR{1}
5750 \POS(-1.5,3.5)*+!DR{2}
5751 \POS(-1,4)*+!DR{3}
5752 \POS(-0.5,4.5)*+!DR{4}
5753 \POS(-0,5)*+!DR{5}
5754 \POS(0,-11.5)*\xybox{
5755 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,5.5)}%
5756 \POS0,{\xylattice{-5}{5}00}%
5757 \def\latticebody{\POS="c"+(-5.5,0)\ar@{--}"c"+(5.5,0)}%
5758 \POS0,{\xylattice00{0}5}%
5759 \POS(0,-0.5)\ar(0,5.5) \POS(0,5.5)*+!UR{p}
5760 \POS(-5.5,0)\ar(5.5,0) \POS(5.5,0)*+!UR{x_1}
5761 \POS(-2,0)*{\bullet}
5762 \POS(-1,0)*{\bullet},*+!DL{1}
5763 \POS(-0,0)*{\bullet},*+!DL{2}
5764 \POS(1,0)*{\bullet},*+!DL{3}
5765 \POS(2,0)*{\bullet},*+!DL{4}
5766 \POS(3,0)*+!DL{5}
5767 \POS(4,0)*+!DL{5}
5768 \POS(5,0)*+!DL{5}
5769 \POS(-2,1)*{\bullet}
5770 \POS(-1,1)*{\bullet},*+!DL{1}
5771 \POS(-0,1)*{\bullet},*+!DL{2}
5772 \POS(1,1)*{\bullet},*+!DL{3}
5773 \POS(2,1)*{\bullet},*+!DL{4}
5774 \POS(3,1)*+!DL{5}
5775 \POS(4,1)*+!DL{5}
5776 \POS(5,1)*+!DL{5}
5777 \POS(-1,2)*{\bullet}
5778 \POS(-0,2)*{\bullet},*+!DL{1}
5779 \POS(1,2)*{\bullet},*+!DL{2}
5780 \POS(2,2)*+!DL{3}
5781 \POS(3,2)*+!DL{3}
5782 \POS(4,2)*+!DL{3}
5783 \POS(5,2)*+!DL{3}
5784 \POS(-1,3)*{\bullet}
5785 \POS(-0,3)*{\bullet},*+!DL{1}
5786 \POS(1,3)*{\bullet},*+!DL{2}
5787 \POS(2,3)*+!DL{3}
5788 \POS(3,3)*+!DL{3}
5789 \POS(4,3)*+!DL{3}
5790 \POS(5,3)*+!DL{3}
5791 \POS(-0,4)*{\bullet}
5792 \POS(1,4)*+!DL{1}
5793 \POS(2,4)*+!DL{1}
5794 \POS(3,4)*+!DL{1}
5795 \POS(4,4)*+!DL{1}
5796 \POS(5,4)*+!DL{1}
5797 \POS(-0,5)*{\bullet}
5798 \POS(1,5)*+!DL{1}
5799 \POS(2,5)*+!DL{1}
5800 \POS(3,5)*+!DL{1}
5801 \POS(4,5)*+!DL{1}
5802 \POS(5,5)*+!DL{1}
5803 \POS(-6,-0.5)\ar(-6,5.5) \POS(-6,5.5)*+!UL{p}
5804 \POS(-6,0)*{\bullet}
5805 \POS(-6,1)*{\bullet}
5806 \POS(-6,2)*{\bullet}
5807 \POS(-6,3)*{\bullet}
5808 \POS(-6,4)*{\bullet}
5809 \POS(-6,5)*{\bullet}
5811 \end{xy}
5812 \caption{The transformed parametric polytope from Example~\ref{ex:projection}
5813 for $0 \le p \le 5$}
5814 \label{f:ex:projection:transformed}
5815 \end{figure}
5817 The first step is to compute the lattice width of $P(p)$.
5818 In Example~\ref{ex:width:2}, we already obtained the decomposition
5819 of the parameter domain into
5821 \begin{aligned}
5822 \hat C_1 &= \{\, p \mid 0 \le p \le 5 \,\} \\
5823 \hat C_2 &= \{\, p \mid -5 \le p \le -1 \,\}
5825 \end{aligned}
5827 with corresponding lattice widths and lattice width directions
5829 \begin{aligned}
5830 \vec c_1 &= (0,1) & w_{\vec c_1}(p) &= 5-p \\
5831 \vec c_2 &= (1,0) & w_{\vec c_2}(p) &= 5+p
5833 \end{aligned}
5835 Note that in this example, the gaps in both coordinate directions
5836 are $1$, so, in principle, there is no need to perform any unimodular
5837 transformation here. Nevertheless, we will apply the transformation
5838 that would be applied by the algorithm.
5839 On the first domain, we extend the lattice width direction
5840 to the unimodular matrix
5842 \begin{bmatrix}
5843 0 & 1 \\
5844 1 & 0
5845 \end{bmatrix}
5848 After application to the existentially quantified variables $\vec x$,
5849 we obtain the parametric polytope
5851 P'_1(p) = \left\{\,
5852 \vec x \mid
5853 \begin{aligned}
5854 -2 x_2 + p + 5 &\ge 0 \\
5855 2 x_2 + p + 5 &\ge 0 \\
5856 -2 x_1 - p + 5 &\ge 0 \\
5857 2 x_1 - p + 5 &\ge 0 \\
5858 p & \ge 0 \\
5859 \end{aligned}
5860 \,\right\}
5862 shown in the top part of \autoref{f:ex:projection:transformed}.
5863 We now temporarily remove the existential quantification on $x_1$,
5864 resulting in
5866 T' = \{ (p, x_1) \in \ZZ^2 \mid \exists x_2 \in \ZZ : (s, \vec x) \in P' \}
5869 Since there is only one existentially quantified variable left,
5870 we know we only have to shift and subtract the set once to obtain
5871 a set with at most one value of $x_2$ associated to each value
5872 of $(p, x_1)$.
5873 In particular, let $f(x,z_1,z_2)$ be the generating function
5874 of the integer points in $P'$. Then $g(x,z_1) = f'(x,z_1,1)$, with
5875 $f'(x,z_1,z_2) = f(x,z_1,z_2) - f(x,z_1,z_2) \star z_2 f(x,z_1,z_2)$,
5876 is the generating function of $T'$.
5878 To check whether we need to subtract any shifted copies of
5879 $g(x,z_1)$ from itself, we compute
5881 h(x,z_1) = g(x,z_1) \star \frac{z_1 \, g(x,z_1)}{1-z_1}
5884 The second argument of this Hadamard product is depicted
5885 in \autoref{f:ex:projection:transformed} by its coefficients.
5886 The exponents in $h(x,z_1)$ that have a non-zero coefficient
5887 are therefore those marked by both a dot ($\bullet$) and
5888 a number. The total sum can be computed as $h(1,1) = 26$,
5889 which is finite, but non-zero. We therefore need to subtract
5890 at least one shifted copy of $g(x,z_1)$.
5893 g'(x,z_1) = g(x,z_1) - g(x,z_1) \star z_1 g(x,z_1)
5896 Computing
5898 h'(x,z_1) = g'(x,z_1) \star \frac{z_1^2 \, g(x,z_1)}{1-z_1}
5901 we would find that $h'(1,1) = 0$ and so we do not need
5902 to shift and subtract any further.
5903 However, since we are dealing with a two-dimensional problem,
5904 we already know from \autoref{l:gap} that we can stop
5905 after one shift and subtract, so we do not even need
5906 to compute $h'(x,z_1)$ here.
5907 The function $g'(x,z_1)$ now only has non-zero coefficients
5908 for at most one exponent of $z_1$ for each exponent of $x$.
5909 We may therefore project down to obtain
5910 the function $g'(x,1)$, which is the generating function
5911 of the set in the lower left part of \autoref{f:ex:projection:transformed}.
5913 For the chamber $\hat C_2$ of the parameter domain, the computations
5914 are nearly identical and the final result is simply the sum
5915 of the two generating functions found for each of the two (disjoint)
5916 chambers.
5918 \end{example}