doc: clean up "exponential substitution" section
[barvinok.git] / doc / implementation.tex
blob91682c27478fc1654c7a332d2fff71da732a6ab6
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 $t = 0$.
1520 Here, we discuss the exponential substitution.
1522 Consider now one of the terms in~\eqref{eq:rgf},
1524 g(t) =
1525 \frac{\sum_{k=1}^{r} e^{a_k t}}
1526 {\prod_{j=1}^{d}\left(1-e^{c_j t}\right)}
1529 with $a_k = \sp{w_{ik}}{\lambda}$ and $c_j = \sp{b_{ij}}{\lambda}$.
1530 We rewrite this equation as
1532 g(t) =
1533 (-1)^d
1534 \frac{\sum_{k=1}^{r} e^{a_k t}}
1535 {t^d \prod_{j=1}^d c_j}
1536 \prod_{j=1}^d \frac{-c_j t}
1537 {1-e^{c_j t}}
1540 The second factor is analytic in a neighborhood of the origin
1541 $t = c_1 = \cdots = c_d = 0$ and therefore has a Taylor series expansion
1542 \begin{equation}
1543 \label{eq:todd}
1544 \prod_{j=1}^d \frac{-c_j t}
1545 {1-e^{c_j t}}
1547 \sum_{m=0}^{\infty} \todd_m(-c_1, \ldots, -c_d) t^m
1549 \end{equation}
1550 where $\todd_m$ is a homogeneous polynomial of degree $m$ called
1551 the $m$-th \ai{Todd polynomial}~\cite{Barvinok1999}.
1552 Also expanding the numerator in the first factor, we find
1554 g(t) = \frac{(-1)^d}{t^d \prod_{j=1}^d c_j}
1555 \left(
1556 \sum_{n=0}^{\infty}\frac{\sum_{k=1}^{r} a_k^n}{n!} t^n
1557 \right)
1558 \left(
1559 \sum_{m=0}^{\infty} \todd_m(-c_1, \ldots, -c_d) t^m
1560 \right)
1563 with constant term
1564 \begin{multline}
1565 \label{eq:todd:constant}
1566 \frac{(-1)^d}{t^d \prod_{j=1}^d c_j}
1567 \left(\sum_{i=0}^d \frac{\sum_{k=1}^{r} a_k^i}{i!}
1568 \todd_{d-i}(-c_1, \ldots, -c_d)\right)t^d
1569 = \\
1570 \frac{(-1)^d}{\prod_{j=1}^d c_j}
1571 \sum_{i=0}^d \frac{\sum_{k=1}^{r} a_k^i}{i!} \todd_{d-i}(-c_1, \ldots, -c_d)
1573 \end{multline}
1574 To compute the first $d+1$ terms in the Taylor series~\eqref{eq:todd},
1575 we write down the truncated Taylor series
1577 \frac{e^t -1}t \equiv
1578 \sum_{i=0}^d \frac 1{(i+1)!} t^i \equiv
1579 \frac 1 {(d+1)!} \sum_{i=0}^d \frac{(d+1)!}{(i+1)!} t^i
1580 \mod t^{d+1}
1583 where we have
1585 \frac 1 {(d+1)!} \sum_{i=0}^d \frac{(d+1)!}{(i+1)!} t^i
1586 \in \frac 1{(d+1)!} \ZZ[t]
1589 Computing the reciprocal as explained in Section~\ref{s:division},
1590 we find
1591 \begin{equation}
1592 \label{eq:t-exp-1}
1593 \frac{-t}{1-e^t} =
1594 \frac{t}{e^t-1} = \frac 1{\frac{e^t -1}t}
1595 \equiv (d+1)! \frac 1{\sum_{i=0}^d \frac{(d+1)!}{(i+1)!} t^i}
1596 \eqqcolon \sum_{i=0}^d b_i t^i
1598 \end{equation}
1599 Note that the constant term of the denominator is $1/(d+1)!$.
1600 The denominators of the quotient are therefore $((d+1)!)^{i+1}/(d+1)!$.
1601 Also note that the $b_i$ are independent of the generating function
1602 and can be computed in advance.
1603 An alternative way of computing the $b_i$ is to note that
1605 \frac{t}{e^t-1} = \sum_{i=0}^\infty B_i \frac{t^i}{i!}
1608 with $B_i = i! \, b_i$ the \ai{Bernoulli number}s, which can be computed
1609 using the recurrence~\eqref{eq:Bernoulli} (see Section~\ref{s:nested}).
1611 Substituting $t$ by $c_j t$ in~\eqref{eq:t-exp-1}, we have
1613 \frac{-c_j t}{1-e^{c_j t}} = \sum_{i=0}^d b_i c_j^i t^i
1616 Multiplication of these truncated Taylor series for each $c_j$
1617 results in the first $d+1$ terms of~\eqref{eq:todd},
1619 \sum_{m=0}^{d} \todd_m(-c_1, \ldots, -c_d) t^m
1620 \eqqcolon
1621 \sum_{m=0}^{d} \frac{\beta_m}{((d+1)!)^m} t^m
1624 from which
1625 it is easy to compute the constant term~\eqref{eq:todd:constant}.
1626 Note that this convolution can also be computed without the use
1627 of rational coefficients,
1629 \frac{(-1)^d}{\prod_{j=1}^d c_j}
1630 \sum_{i=0}^d \frac{\alpha_i}{i!} \frac{\beta_{d-i}}{((d+1)!)^{d-i}}
1632 \frac{(-1)^d}{((d+1)!)^d\prod_{j=1}^d c_j}
1633 \sum_{i=0}^d \left(\frac{((d+1)!)^i}{i!}\alpha_i\right) \beta_{d-i}
1636 with $\alpha_i = \sum_{k=1}^{r} a_k^i$.
1638 \begin{example}
1639 Consider the \rgf/
1640 \begin{multline*}
1641 \f T x =
1642 \frac{x_1^2}{(1-x_1^{-1})(1-x_1^{-1}x_2)}
1644 \frac{x_2^2}{(1-x_2^{-1})(1-x_1 x_2^{-1})}
1645 + {} \\
1646 \frac1{(1-x_1)(1-x_2)}
1647 \end{multline*}
1648 from \shortciteN[Example~39]{Verdoolaege2005PhD}.
1649 Since this is a 2-dimensional problem, we first compute the first
1650 3 Todd polynomials (evaluated at $-1$),
1652 \frac{e^t -1}t \equiv
1653 1 + \frac 1 2 t + \frac 1 6 t^2 =
1654 \frac 1 6
1655 \begin{bmatrix}
1656 6 & 3 & 1
1657 \end{bmatrix}
1661 \frac {-t}{1-e^t} =
1662 \frac t{e^t -1} \equiv
1663 \begin{bmatrix}
1664 \displaystyle\frac{1}{1} & \displaystyle\frac{-3}{6} & \displaystyle\frac{3}{36}
1665 \end{bmatrix}
1668 where we represent each truncated power series by a vector of its
1669 coefficients.
1670 The vector $\vec\lambda = (1, -1)$ is not
1671 orthogonal to any of the rays, so we can use the substitution
1672 $\vec x = e^{(1, -1)t}$
1673 and obtain
1675 \frac{e^{2t}}{(1-e^{-t})(1-e^{-2t})}
1677 \frac{e^{-2t}}{(1-e^{t})(1-e^{2t})}
1679 \frac1{(1-e^{t})(1-e^{-t})}
1682 We have
1683 \begin{align*}
1684 \frac{t}{1-e^{- t}} & =
1685 \begin{bmatrix}
1686 \displaystyle\frac{1}{1} & \displaystyle\frac{3}{6} & \displaystyle\frac{3}{36}
1687 \end{bmatrix}
1689 \frac{2t}{1-e^{-2 t}} & =
1690 \begin{bmatrix}
1691 \displaystyle\frac{1}{1} & \displaystyle\frac{6}{6} & \displaystyle\frac{12}{36}
1692 \end{bmatrix}
1694 \frac{-t}{1-e^{t}} & =
1695 \begin{bmatrix}
1696 \displaystyle\frac{1}{1} & \displaystyle\frac{-3}{6} & \displaystyle\frac{3}{36}
1697 \end{bmatrix}
1699 \frac{-2t}{1-e^{2t}} & =
1700 \begin{bmatrix}
1701 \displaystyle\frac{1}{1} & \displaystyle\frac{-6}{6} & \displaystyle\frac{12}{36}
1702 \end{bmatrix}
1704 \end{align*}
1705 The first term in the \rgf/ evaluates to
1706 \begin{align*}
1708 \frac 1{-1 \cdot -2}
1709 \begin{bmatrix}
1710 \displaystyle\frac{1}{1} & \displaystyle\frac{2}{1} & \displaystyle\frac{4}{2}
1711 \end{bmatrix}
1713 \left(
1714 \begin{bmatrix}
1715 \displaystyle\frac{1}{1} & \displaystyle\frac{3}{6} & \displaystyle\frac{3}{36}
1716 \end{bmatrix}
1717 \begin{bmatrix}
1718 \displaystyle\frac{1}{1} & \displaystyle\frac{6}{6} & \displaystyle\frac{12}{36}
1719 \end{bmatrix}
1720 \right)
1722 = {} &
1723 \frac 1{2}
1724 \begin{bmatrix}
1725 \displaystyle\frac{1}{1} & \displaystyle\frac{2}{1} & \displaystyle\frac{4}{2}
1726 \end{bmatrix}
1728 \begin{bmatrix}
1729 \displaystyle\frac{1}{1} & \displaystyle\frac{9}{6} & \displaystyle\frac{33}{36}
1730 \end{bmatrix}
1732 = {} &
1733 \frac 1{72}
1734 \begin{bmatrix}
1735 1 & 2 \cdot 6 & 4 \cdot 18
1736 \end{bmatrix}
1738 \begin{bmatrix}
1739 1 & 9 & 33
1740 \end{bmatrix}
1741 = \frac {213}{72} = \frac{71}{24}
1743 \end{align*}
1744 Due to symmetry, the second term evaluates to the same value,
1745 while for the third term we find
1747 \frac{1}{-1\cdot 1 \cdot 36}
1748 \begin{bmatrix}
1749 1 & 0 \cdot 6 & 0 \cdot 18
1750 \end{bmatrix}
1752 \begin{bmatrix}
1753 1 & 0 & -3
1754 \end{bmatrix}
1756 \frac{-3}{-36} = \frac 1{12}
1759 The sum is
1761 \frac{71}{24} + \frac{71}{24} + \frac 1{12} = 6
1764 \end{example}
1766 Note that the run-time complexities of polynomial and exponential
1767 substitution are basically the same. The experiments of
1768 \citeN{Koeppe2006primal} are somewhat misleading in this respect
1769 since the polynomial substitution (unlike the exponential
1770 substitution) had not been optimized to take full
1771 advantage of the stopped Barvinok decomposition.
1772 For comparison, \autoref{t:hickerson} shows running times
1773 for the same experiments of that paper, but using
1774 barvinok version \verb+barvinok-0.23-47-gaa9024e+
1775 on an Athlon MP 1500+ with 512MiB internal memory.
1776 This machine appears to be slightly slower than the
1777 machine used in the experiments of \citeN{Koeppe2006primal}
1778 as computing {\tt hickerson-14} using the dual decomposition
1779 with polynomial substitution and maximal index 1
1780 took 2768 seconds on this machine using \LattEmk/.
1781 At this stage, it is not clear yet why the number of
1782 cones in the dual decomposition of {\tt hickerson-13}
1783 differs from that of \LattE/~\shortcite{latte1.1} and
1784 \LattEmk/~\cite{latte-macchiato}.
1785 We conclude from \autoref{t:hickerson} that (our implementation of)
1786 the exponential substitution is always slightly faster than
1787 (our implementation of) the polynomial substitution.
1788 The optimal maximal index for these examples is about 500,
1789 which agrees with the experiments of \citeN{Koeppe2006primal}.
1791 \begin{table}
1792 \begin{center}
1793 \begin{tabular}{rrrrrrr}
1794 \hline
1796 \multicolumn{3}{c}{Dual decomposition} &
1797 \multicolumn{3}{c}{Primal decomposition}
1800 & \multicolumn{2}{c}{Time (s)} &
1801 & \multicolumn{2}{c}{Time (s)}
1803 \cline{3-4}
1804 \cline{6-7}
1805 Max.\ index & Cones & Poly & Exp & Cones & Poly & Exp \\
1806 \hline
1807 \multicolumn{7}{c}{{\tt hickerson-12}}
1809 \hline
1810 1 & 11625 & 9.24 & 8.90 & 7929 & 4.80 & 4.55
1812 10 & 4251 & 4.32 & 4.19 & 803 & 0.66 & 0.62
1814 100 & 980 & 1.42 & 1.35 & 84 & 0.13 & 0.12
1816 200 & 550 & 1.00 & 0.92 & 76 & 0.12 & 0.12
1818 300 & 474 & 0.93 & 0.86 & 58 & 0.12 & 0.10
1820 500 & 410 & 0.90 & 0.83 & 42 & 0.10 & 0.10
1822 1000 & 130 & 0.42 & 0.38 & 22 & {\bf 0.10} & {\bf 0.07}
1824 2000 & 10 & {\bf 0.10} & {\bf 0.10} & 22 & 0.10 & 0.09
1826 5000 & 7 & 0.12 & 0.11 & 7 & 0.12 & 0.10
1828 \hline
1829 \multicolumn{7}{c}{{\tt hickerson-13}}
1831 \hline
1832 1 & 494836 & 489 & 463 & 483507 & 339 & 315
1834 10 & 296151 & 325 & 309 & 55643 & 51 & 48
1836 100 & 158929 & 203 & 192 & 9158 & 11 & 10
1838 200 & 138296 & 184 & 173 & 6150 & 9 & 8
1840 300 & 110438 & 168 & 157 & 4674 & 8 & 7
1842 500 & 102403 & 163 & 151 & 3381 & {\bf 8} & {\bf 7}
1844 1000 & 83421 & {\bf 163} & {\bf 149} & 2490 & 8 & 7
1846 2000 & 77055 & 170 & 153 & 1857 & 10 & 8
1848 5000 & 57265 & 246 & 211 & 1488 & 13 & 11
1850 10000 & 50963 & 319 & 269 & 1011 & 26 & 21
1852 \hline
1853 \multicolumn{7}{c}{{\tt hickerson-14}}
1855 \hline
1856 1 & 1682743 & 2171 & 2064 & 552065 & 508 & 475
1858 10 & 1027619 & 1453 & 1385 & 49632 & 62 & 59
1860 100 & 455474 & 768 & 730 & 8470 & 14 & 13
1862 200 & 406491 & 699 & 661 & 5554 & 11 & 10
1864 300 & 328340 & 627 & 590 & 4332 & 11 & 9
1866 500 & 303566 & 605 & 565 & 3464 & {\bf 11} & {\bf 9}
1868 1000 & 232626 & {\bf 581} & {\bf 532} & 2384 & 12 & 10
1870 2000 & 195368 & 607 & 545 & 1792 & 14 & 12
1872 5000 & 147496 & 785 & 682 & 1276 & 19 & 16
1874 10000 & 128372 & 966 & 824 & 956 & 29 & 23
1876 \hline
1877 \end{tabular}
1878 \caption{Timing results of dual and primal decomposition with
1879 polynomial or exponential substitution on the Hickerson examples}
1880 \label{t:hickerson}
1881 \end{center}
1882 \end{table}
1884 \subsection{Approximate Enumeration using Nested Sums}
1885 \label{s:nested}
1887 If $P \in \QQ^d$ is a polyhedron and $p(\vec x) \in \QQ[\vec x]$ is a
1888 polynomial and we want to sum $p(\vec x)$ over all integer values
1889 of (a subset of) the variables $\vec x$, then we can do this incrementally
1890 by taking a variable $x_1$ with lower bound $L(\vec{\hat x})$
1891 and upper bound $U(\vec{\hat x})$, with $\vec{\hat x} = (x_2, \ldots, x_d)$,
1892 and computing
1893 \begin{equation}
1894 \label{eq:nested:sum}
1895 Q(\vec{\hat x}) = \sum_{x_1 = L(\vec{\hat x})}^{U(\vec{\hat x})} p(\vec x)
1897 \end{equation}
1898 Since $P$ is a polytope, the lower bound is a maximum of affine expressions
1899 in the remaining variables, while the upper bound is a minimum of such expressions.
1900 If the coefficients in these expressions are all integer, then we can
1901 compute $Q(\vec{\hat x})$ exactly as a piecewise polynomial using formulas
1902 for sums of powers, as proposed by, e.g.,
1903 \shortciteN{Tawbi1994,Sakellariou1997sums,VanEngelen2004}.
1904 If some of the coefficients are not integer, we can apply the same formulas
1905 to obtain an approximation, which can is some cases be shown
1906 to be an overapproximation~\shortcite{VanEngelen2004}.
1907 Note that if we take the initial polynomial to be the constant $1$, then
1908 this gives us a method for computing an approximation of the number
1909 of integer points in a (parametric) polytope.
1911 The first step is to compute the chamber decomposition of $P$ when viewed
1912 as a 1-dimensional parametric polytope. That is, we need to partition
1913 the projection of $P$ onto the remaining variables into polyhedral cells
1914 such that in each cell, both the upper and the lower bound are described
1915 by a single affine expression. Basically, for each pair of lower and upper
1916 bound, we compute the cell where the chosen lower bound is (strictly)
1917 smaller than all other lower bounds and similarly for the upper bound.
1919 For any given pair of lower and upper bound $(l(\vec {\hat x}), u(\vec{\hat x}))$,
1920 the formula~\eqref{eq:nested:sum} is computed for each monomial of $p(\vec x)$
1921 separately. For the constant term $\alpha_0$, we have
1922 \begin{equation}
1923 \label{eq:summation:1d}
1924 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_0(\vec{\hat x})
1925 = \alpha_0(\vec{\hat x}) \left(u(\vec{\hat x}) - l(\vec {\hat x}) + 1\right)
1927 \end{equation}
1928 For the higher degree monomials, we use the formula
1929 \begin{equation}
1930 \label{eq:summation}
1931 \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}
1932 =: S_n(m)
1934 \end{equation}
1935 with $B_i$ the \ai{Bernoulli number}s, which can be computed
1936 using the recurrence
1937 \begin{equation}
1938 \label{eq:Bernoulli}
1939 \sum_{j=0}^m{m+1\choose{j}}B_j = 0
1940 \qquad B_0 = 1
1942 \end{equation}
1943 Note that \eqref{eq:summation} is also valid if $m = 0$,
1944 i.e., $S_n(0) = 0$, a fact
1945 that can be easily shown using Newton series~\shortcite{VanEngelen2004}.
1947 \newcounter{saveenumi}
1949 Since we can only directly apply the summation formula when
1950 the lower bound is zero (or one), we need to consider several
1951 cases.
1952 \begin{enumerate}
1953 \item $l(\vec {\hat x}) \ge 1$
1954 \label{i:l}
1956 \begin{align*}
1957 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1959 \alpha_n(\vec{\hat x})
1960 \left(
1961 \sum_{x_1 = 1}^{u(\vec{\hat x})} x_1^n
1963 \sum_{x_1 = 1}^{l(\vec {\hat x})-1} x_1^n
1964 \right)
1967 \alpha_n(\vec{\hat x})
1968 \left( S_n(u(\vec{\hat x})+1) - S_n(l(\vec {\hat x})) \right)
1969 \end{align*}
1971 \item $u(\vec{\hat x}) \le -1$
1972 \label{i:u}
1974 \begin{align*}
1975 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1977 \alpha_n(\vec{\hat x}) (-1)^n
1978 \sum_{x_1 = -u(\vec {\hat x})}^{-l(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1981 \alpha_n(\vec{\hat x}) (-1)^n
1982 \left( S_n(-l(\vec{\hat x})+1) - S_n(-u(\vec {\hat x})) \right)
1983 \end{align*}
1985 \item $l(\vec {\hat x}) \le 0$ and $u(\vec{\hat x}) \ge 0$
1986 \label{i:l:u}
1988 \begin{align*}
1989 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
1991 \alpha_n(\vec{\hat x})
1992 \left(
1993 \sum_{x_1 = 0}^{u(\vec{\hat x})} x_1^n
1995 (-1)^n
1996 \sum_{x_1 = 1}^{-l(\vec {\hat x})} x_1^n
1997 \right)
2000 \alpha_n(\vec{\hat x})
2001 \left(
2002 S_n(u(\vec{\hat x})+1)
2004 (-1)^n
2005 S_n(-l(\vec{\hat x})+1)
2006 \right)
2007 \end{align*}
2009 \setcounter{saveenumi}{\value{enumi}}
2010 \end{enumerate}
2012 If the coefficients in the lower and upper bound are all
2013 integer, then the above 3 cases partition (the integer points in)
2014 the projection of $P$ onto the remaining variables.
2015 However, if some of the coefficients are rational, then the lower
2016 and upper bound can lie in the open interval $(0,1)$ for some
2017 values of $\vec{\hat x}$. We may therefore also want to consider
2018 the following two cases.
2020 \begin{enumerate}
2021 \setcounter{enumi}{\value{saveenumi}}
2022 \item $0 < l(\vec {\hat x}) < 1$
2023 \label{i:l:fractional}
2026 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
2028 \alpha_n(\vec{\hat x})
2029 S_n(u(\vec{\hat x})+1)
2032 \item $0 < -u(\vec {\hat x}) < 1$
2033 \label{i:u:fractional}
2036 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_n(\vec{\hat x}) \, x_1^n
2038 \alpha_n(\vec{\hat x})
2039 (-1)^n
2040 S_n(-l(\vec{\hat x})+1)
2043 \end{enumerate}
2044 Note that we may add the constraint $u \ge 1$ to
2045 case~\ref{i:l:fractional} and the constraint $l \le -1$
2046 to case~\ref{i:u:fractional}, since the correct value for
2047 these two cases would be zero if these extra constraints do not hold.
2049 An alternative to adding the above two cases would be
2050 to simply ignore them, i.e., assume a value of $0$.
2051 Another alternative would be to reduce case~\ref{i:l:u}
2054 l(\vec {\hat x}) \le -1\quad\hbox{and}\quad u(\vec{\hat x}) \ge 1
2057 while extending cases~\ref{i:l:fractional} and~\ref{i:u:fractional}
2060 -1 < l(\vec {\hat x}) < 1\quad\hbox{and}\quad u \ge 1
2064 -1 < u(\vec {\hat x}) < 1\quad\hbox{and}\quad l \le -1
2067 respectively, with the remaining cases
2068 ($-1 < l \le u < 1$) having value $0$.
2069 There does not appear to be a consistently better choice
2070 here, as each of these three approaches seems to yield better
2071 results on some examples.
2072 The last approach has the additional drawback that we
2073 would also have to deal with 5 cases, even if the bounds
2074 are integers.
2076 If at least one of the lower or upper bound is an
2077 integer affine expression, then we can reduce
2078 the 3 (or 5) cases to a single case (case~\ref{i:l:u})
2079 by an affine substitution that ensure that the
2080 new (lower or upper) bound is zero.
2081 In particular, if $l(\vec {\hat x})$ is an integer affine
2082 expression, then we replace $x$ by $x' + l(\vec {\hat x})$
2083 and similarly for an upper bound.
2085 \subsection{Exact Enumeration using Nested Sums}
2086 \label{s:nested:exact}
2088 The exact enumeration using nested sums proceeds in much
2089 the same way as the approximate enumeration from
2090 \autoref{s:nested}, with the notable exception
2091 that we need to take the (greatest or least) integer part
2092 of any fractional bounds that may occur.
2093 This has several consequences, discussed below.
2095 Since we will introduce floors during the recursive application
2096 of the procedure, we may as well allow the weight
2097 $p(\vec x)$ in~\eqref{eq:nested:sum} to be a (piecewise)
2098 quasipolynomial.
2100 For the constant term, \eqref{eq:summation:1d} becomes
2102 \sum_{x_1 = l(\vec {\hat x})}^{u(\vec{\hat x})} \alpha_0(\vec{\hat x})
2103 = \alpha_0(\vec{\hat x})
2104 \left(\floor{u(\vec{\hat x})} - \ceil{l(\vec {\hat x})} + 1\right)
2108 Since we force the lower and upper bounds to be integers,
2109 cases~\ref{i:l:fractional} and~\ref{i:u:fractional} do not occur,
2110 while the conditions for cases~\ref{i:l} and~\ref{i:u} can be simplified
2113 l(\vec {\hat x}) > 0
2117 u(\vec {\hat x}) < 0
2120 respectively.
2122 If the variable $x$ appears in any floor expression, either
2123 because such an expression was present in the original weight function
2124 or because it was introduced when another variable with an affine bound
2125 in $x$ was summed, then the domain has to
2126 be ``splintered'' into $D$ parts, where $D$ is the least common
2127 multiple of the denominators of the coefficients of $x$ in
2128 any of the integer parts.
2129 In particular, the domain is split into $x = D y + i$ for
2130 each $i$ in $[0, D-1]$. Since $D$ is proportional to
2131 the coefficients in the constraints, it is exponential in
2132 the input size. This splintering will therefore
2133 introduce exponential behavior, even if the dimension is fixed.
2135 Splintering is clearly the most expensive step in the algorithm,
2136 so we want to avoid this step as much as possible.
2137 \shortciteN{Pugh94counting} already noted that summation should
2138 proceed over variables with integer bounds first.
2139 This can be extended to choosing a variable with the smallest
2140 coefficient in absolute value. In this way, we can avoid
2141 splintering on the largest denominator.
2143 \shortciteN{Sakellariou1996phd} claims that splintering
2144 can be avoided altogether.
2145 In particular, \shortciteN[Lemma~3.2]{Sakellariou1996phd}
2146 shows that
2148 \sum_{x=0}^a x^m\left(x \bmod b\right)^n
2151 with $a$ and $b$ integers, is equal to
2152 \begin{equation}
2153 \label{eq:3.2}
2154 \begin{cases}
2155 \displaystyle
2156 \sum_{x=0}^a x^{m+n} & \text{if $a<b$}
2158 \displaystyle
2159 \sum_{i=0}^{\floor{a/b}-1} \sum_{x=0}^{b-1} (x+ib)^m x^n +
2160 \sum_{x=0}^{a \bmod b} (x+b\floor{a/b})^m x^n & \text{if $a\ge b$}
2162 \end{cases}
2163 \end{equation}
2164 effectively avoiding splintering if a given monomial contains
2165 a single integer part expression with argument of the form
2166 $x/b$. An argument of the form $(x-c(\vec{\hat x}))/b$ can
2167 be handled through a variable substitution.
2168 If the argument is of the form $c x/b$, with $c \ne 1$,
2169 then \shortciteN[(3.27)]{Sakellariou1996phd} proposes to
2170 rewrite the monomial as
2171 \begin{align*}
2172 \sum_{x=0}^a (c x \bmod b)^n
2174 \sum_{x=0}^a \sum_{y=cx}^{cx} (y \bmod b)^n
2177 \sum_{x=0}^a
2178 \left(\sum_{y=0}^{cx} (y \bmod b)^n - \sum_{y=0}^{cx-1} (y \bmod b)^n\right)
2179 \end{align*}
2180 and applying \eqref{eq:3.2}.
2181 However, such an application results in an expression containing
2183 \sum_{y=0}^{cx \bmod b} y^n
2186 which in turn leads to a polynomial of degree $n+1$ in $(c x \bmod b)$,
2187 i.e., of degree one higher than the original expression.
2188 Furthermore, if the bound on $x$ is rational then $a$ itself contains
2189 a floor, which, on application of \eqref{eq:3.2}, results in
2190 a nested floor expression, blocking the application of the same
2191 rule for the next variable.
2192 Finally, the case where a monomial contains multiple floor
2193 expressions, either occurring in the input quasi-polynomial
2194 or introduced by different variables having a rational
2195 bound with a non-zero coefficient in the same variable, is not handled.
2196 Also note that if we disallow nested floor expressions,
2197 then this rule will rarely be applicable since we try to eliminate
2198 variables with integer bounds first.
2200 \subsection{Summation using local Euler-Maclaurin formula}
2201 \label{s:euler}
2203 \sindex{local}{Euler-Maclaurin formula}
2204 In this section we provide some implementation details
2205 on using \ai{local Euler-Maclaurin formula} to compute
2206 the sum of a piecewise polynomial evaluated in all integer
2207 points of a two-dimensional parametric polytope.
2208 For the theory behind these formula and a discussion
2209 of the original implementation (for non-parametric simplices),
2210 we refer to \shortciteN{Berline2006local}.
2212 In particular, consider a parametric piecewise polynomial
2213 in $n$ parameters and $m$ variables
2214 $c : \ZZ^n \to \ZZ^m \to \QQ : \vec p \mapsto c(\vec p)$,
2215 with $c(\vec p) : \ZZ^m \to \QQ : \vec x \mapsto c(\vec p)(\vec x)$
2218 c_{\vec p}(\vec x) =
2219 \begin{cases}
2220 c_1(\vec p)(\vec x) & \text{if $\vec x \in D_1(\vec p)$}
2222 \vdots
2224 c_r(\vec p)(\vec x) & \text{if $\vec x \in D_r(\vec p)$}
2226 \end{cases}
2228 with the $c_i$ polynomials, $c_i \in (\QQ[\vec p])[\vec x]$, and
2229 the $D_i$ disjoint linearly parametric polytopes.
2230 We want to compute
2232 g(\vec p) = \sum_{\vec x \in \ZZ^m} c(\vec p)(\vec x)
2236 \subsubsection{Reduction to the summation of a parametric polynomial
2237 over a parametric polytope with a fixed combinatorial structure}
2239 Since the $D_i$ are disjoint, we can consider each
2240 $(c_i, D_i)$-pair individually and compute
2242 g(\vec p) = \sum_{i=1}^r g_i(\vec p) =
2243 \sum_{i=1}^r \sum_{\vec x \in D_r(\vec p) \cap \ZZ^m} c_r(\vec p)(\vec x)
2246 The second step is to compute the \ai{chamber decomposition}
2247 ~\shortcite[Section 4.2.3]{Verdoolaege2005PhD} of each parametric
2248 polytope $D_i$.
2249 The result is a subdivision of the parameter space into chambers
2250 $C_{ij}$ such that $D_i$ has a fixed combinatorial structure,
2251 in particular a fixed set of parametric vertices,
2252 on (the interior of) each $C_{ij}$. Applying \autoref{p:inclusion-exclusion},
2253 this subdivision can be transformed into a partition
2254 $\{\, \tilde C_{ij} \,\}$ by
2255 making some of the facets of the chambers open%
2256 ~\shortcite[Section~3.2]{Koeppe2008parametric}.
2257 Since we are only interested in integer parameter values,
2258 any of the resulting open facets $\sp a p + c > 0$,
2259 with $\vec a \in \ZZ^n$ and $c \in \ZZ$,
2260 can then be replaced by $\sp a p + c-1 \ge 0$.
2261 Again, we have
2263 g_i(\vec p) = \sum_j g_{ij}(\vec p) =
2264 \sum_j \sum_{\vec x \in C_{ij}(\vec p) \cap \ZZ^m} c_r(\vec p)(\vec x)
2268 After this reduction, the technique of
2269 \shortciteN{Berline2006local} can be applied practically verbatim
2270 to the parametric polytope with a fixed combinatorial structure.
2271 In principle, we could also handle piecewise quasi-polynomials
2272 using the technique of \shortciteN[Section~4.5.4]{Verdoolaege2005PhD},
2273 except that we only need to create an extra variable for each
2274 distinct floor expression in a monomial, rather than for each
2275 occurrence of a floor expression in a monomial.
2276 However, since we currently only support two-dimensional polytopes,
2277 this reduction has not been implemented yet.
2279 \subsubsection{Summation over a one-dimensional parametric polytope}
2281 The basis for the summation technique is the local
2282 Euler-Maclaurin formula~\cite[Theorem~26]{Berline2006local}
2283 \begin{equation}
2284 \label{eq:EML}
2285 \sum_{\vec x \in P(\vec p) \cap \Lambda} h(\vec p)(\vec x)
2286 = \sum_{F(\vec p) \in {\mathcal F}(P(\vec p))}
2287 \int_{F(\vec p)} D_{P(\vec p),F(\vec p)} \cdot h(\vec p)
2289 \end{equation}
2290 where $P(\vec p)$ is a parametric polytope,
2291 $\Lambda$ is a lattice, ${\mathcal F}(P(\vec p))$
2292 are the faces of $P(\vec p)$, $D_{P(\vec p),F(\vec p)}$ is a
2293 specific differential operator associated to the face of a polytope.
2294 The \ai{Lebesgue measure} used in the integral is such that the
2295 integral of the indicator function of a lattice element of
2296 the lattice $\Lambda \cap (\affhull(F(\vec p)) - F(\vec p))$ is 1,
2297 i.e., the intersection of $\Lambda$ with the linear subspace
2298 parallel to the affine hull of the face $F(\vec p)$.
2299 Note that the original theorem is formulated for a non-parametric
2300 polytope and a non-parametric polynomial. However, as we will see,
2301 in each of the steps in the computation, the parameters can be
2302 treated as symbolic constants without affecting the validity of the formula,
2303 see also~\shortciteN[Section 6]{Berline2006local}.
2305 The differential operator $D_{P(\vec p),F(\vec p)}$ is obtained
2306 by plugging in the vector $\vec D=(D_1,\ldots,D_m)$ of first
2307 order differential operators, i.e., $D_k$ is the first order
2308 differential operator in the $k$th variable,
2309 in the function $\mu_{P(\vec p),F(\vec p)}$.
2310 This function is determined by the \defindex{transverse cone}
2311 of the polyhedron $P(\vec p)$ along its face $F(\vec p)$,
2312 which is the \ai{supporting cone} of $P(\vec p)$ along $F(\vec p)$
2313 projected into the linear subspace orthogonal to $F(\vec p)$.
2314 The lattice associated to this space is the projection of
2315 $\Lambda$ into this space.
2317 In particular, for a zero-dimensional affine cone in the zero-dimensional
2318 space, we have $\mu = 1$~\cite[Proposition 12]{Berline2006local},
2319 while for a one-dimensional affine
2320 cone $K = (-t + \RR_+) r$ in the one-dimensional space, where
2321 $r$ is a primitive integer vector and $t \in [0,1)$,
2322 we have~\cite[(13)]{Berline2006local}
2323 \begin{equation}
2324 \label{eq:mu:1}
2325 \mu(K)(\xi) = \frac{e^{t y}}{1-e^y} + \frac 1{y}
2326 = -\sum_{n=0}^\infty \frac{b(n+1, t)}{(n+1)!} y^n
2328 \end{equation}
2329 with $y = \sps \xi r$ and $b(n,t)$ the \ai{Bernoulli polynomial}s
2330 defined by the generating series
2332 \frac{e^{ty} y}{e^y - 1} = \sum_{n=0}^\infty \frac{b(n,t)}{n!} y^n
2335 The constant terms of these Bernoulli polynomials
2336 are the \ai{Bernoulli number}s.
2338 Applying \eqref{eq:EML} to a one-dimensional parametric polytope
2339 $P(\vec p) = [v_1(\vec p), v_2(\vec p)]$, we find
2341 \begin{aligned}
2342 \sum_{x \in P(\vec p) \cap \ZZ} h(\vec p)(x)
2343 = & \int_{P(\vec p)} D_{P(\vec p), P(\vec p)} \cdot h(\vec p)
2345 & + \int_{v_1(\vec p)} D_{P(\vec p), v_1(\vec p)} \cdot h(\vec p)
2347 & + \int_{v_2(\vec p)} D_{P(\vec p), v_2(\vec p)} \cdot h(\vec p)
2349 \end{aligned}
2351 The transverse cone of a polytope along the whole polytope is
2352 a zero-dimensional cone in a zero-dimensional space and so
2353 $D_{P(\vec p), P(\vec p)} = \mu_{P(\vec p), P(\vec p)}(D) = 1$.
2354 The transverse cone along $v_1(\vec p)$ is $v_1(\vec p) + \RR_+$
2355 and so $D_{P(\vec p), v_1(\vec p)} = \mu(v_1(\vec p) + \RR_+)(D)$
2356 as in \eqref{eq:mu:1}, with $y = \sps D 1 = D$ and
2357 $t = \ceil{v_1(\vec p)} - v_1(\vec p) =
2358 \fractional{-v_1(\vec p)}$.
2359 Similarly we find
2360 $D_{P(\vec p), v_2(\vec p)} = \mu(v_2(\vec p) - \RR_+)(D)$
2361 as in \eqref{eq:mu:1}, with $y = \sps D {-1} = -D$ and
2362 $t = v_2(\vec p) - \floor{v_2(\vec p)} =
2363 \fractional{v_2(\vec p)}$.
2364 Summarizing, we find
2366 \begin{aligned}
2367 \sum_{x \in P(\vec p) \cap \ZZ} h(\vec p)(x)
2368 = & \int_{v_1(\vec p)}^{v_2(\vec p)} h(\vec p)(t) \, dt
2370 & -\sum_{n=0}^\infty \frac{b(n+1, \fractional{-v_1(\vec p)})}{(n+1)!}
2371 (D^n h(\vec p))(v_1(\vec p))
2373 & -\sum_{n=0}^\infty (-1)^n \frac{b(n+1, \fractional{v_2(\vec p)})}{(n+1)!}
2374 (D^n h(\vec p))(v_2(\vec p))
2376 \end{aligned}
2379 Note that in order to apply this formula, we need to verify
2380 first that $v_1(\vec p)$ is indeed smaller than (or equal to)
2381 $v_2(\vec p)$. Since the combinatorial structure of $P(\vec p)$
2382 does not change throughout the interior of the chamber, we only
2383 need to check the order of the two vertices for one value
2384 of the parameters from the interior of the chamber, a point
2385 which we may compute as in \autoref{s:interior}.
2387 \subsubsection{Summation over a two-dimensional parametric polytope}
2389 For two-dimensional polytope, formula~\eqref{eq:EML} has three kinds
2390 of contributions: the integral of the polynomial over the polytope,
2391 contributions along edges and contributions along vertices.
2392 As suggested by~\citeN{Berline2007personal}, the integral can be computed
2393 by applying the Green-Stokes theorem:
2395 \iint_{P(\vec p)}
2396 \left(\frac{\partial M}{\partial x} - \frac{\partial L}{\partial y}\right) =
2397 \int_{\partial P(\vec p)} (L\, dx + M\, dy)
2400 In particular, if $M(\vec p)(x,y)$ is such that
2401 $\frac{\partial M}{\partial x}(\vec p)(x,y) = h(\vec p)(x,y)$
2402 then
2404 \iint_{P(\vec p)} h(\vec p)(x,y) =
2405 \int_{\partial P(\vec p)} M(\vec p)(x,y) \, dy
2408 Care must be taken to integrate over the boundary in the positive
2409 direction. Assuming the vertices of the polygon are not given
2410 in a predetermined order, we can check the correct orientation
2411 of the vertices of each edge individually. Let $\vec n = (n_1, n_2)$
2412 be the inner normal of a facet and let $\vec v_1(\vec p)$
2413 and $\vec v_2(\vec p)$ be the two vertices of the facet, then
2414 the vertices are in the correct order if
2416 \begin{vmatrix}
2417 v_{2,1}(\vec p)-v_{1,1}(\vec p) & n_1
2419 v_{2,2}(\vec p)-v_{1,2}(\vec p) & n_2
2420 \end{vmatrix}
2421 \ge 0
2424 Since these two vertices belong to the same edge, their order
2425 will not change within a chamber and so we can again perform
2426 this check for a single value of the parameters.
2427 To integrate $M$ over an edge $F$, let $\vec f$ be a primitive
2428 integer vector in the direction of the edge.
2429 Then $\vec v_2(\vec p) = \vec v_1(\vec p) + k(\vec p) \, \vec f$
2430 and any point on the edge can be written as
2431 $\vec v_1(\vec p) + \lambda \vec f$ with
2432 $0 \le \lambda \le k(\vec p)$.
2433 That is,
2434 \begin{equation}
2435 \label{eq:EML:int}
2436 \int_F M(\vec p)(x,y) \, dy
2438 \int_0^{k(\vec p)}
2439 M(\vec p)(v_{1,1}(\vec p) + \lambda f_1,
2440 v_{1,2}(\vec p) + \lambda f_2)
2441 f_2 \, d\lambda
2443 \end{equation}
2445 For the edges, we can again apply \eqref{eq:mu:1}, but we
2446 must first project the supporting cone at the edge into
2447 the linear subspace orthogonal to the edge.
2448 Let $\vec n = (n_1, n_2)$ be the (primitive integer) inner normal
2449 of this facet $F(\vec p)$, then $\vec f = (-n_2, n_1)$ is parallel
2450 to the facet and we can write one of the vertices $\vec v(\vec p)$
2451 as a linear combination of these two vectors:
2452 \begin{equation}
2453 \label{eq:EML:facet:coordinates}
2454 \vec v(\vec p)
2456 \begin{bmatrix}
2457 \vec f & \vec n
2458 \end{bmatrix}
2459 \vec a(\vec p)
2461 \begin{bmatrix}
2462 -n_2 & n_1 \\
2463 n_1 & n_2
2464 \end{bmatrix}
2465 \vec a(\vec p)
2466 \end{equation}
2468 \begin{equation}
2469 \label{eq:EML:facet:coordinates:2}
2470 \vec a(\vec p)
2472 \begin{bmatrix}
2473 -n_2 & n_1 \\
2474 n_1 & n_2
2475 \end{bmatrix}^{-1}
2476 \vec v(\vec p)
2478 \begin{bmatrix}
2479 -n_2/d & n_1/d \\
2480 n_1/d & n_2/d
2481 \end{bmatrix}
2482 \vec v(\vec p),
2483 \end{equation}
2484 with $d = n_1^2+n_2^2$.
2485 The lattice associated to the linear subspace orthogonal
2486 to the facet is the projection of $\Lambda$ into this space.
2487 Since $\vec n$ is primitive, a basis for this lattice can be
2488 identified with $\vec n/d$.
2489 The coordinate of the whole facet in this space is therefore
2491 d a_2(\vec p) =
2492 \begin{bmatrix}
2493 n_1 & n_2
2494 \end{bmatrix}
2495 \vec v(\vec p)
2496 $, while the transverse cone is $d a_2(\vec p) + \RR_+$.
2497 Similarly, a linear functional $\vec \xi'$ projects onto
2498 a linear functional $\xi = \sp {\xi'} n/d$ in the linear subspace.
2499 Applying \eqref{eq:mu:1}, with $y = \frac{n_1}d D_1 + \frac{n_2}d D_2$
2500 and $t = \fractional{- n_1 v_1(\vec p) - n_2 v_2(\vec p)}$, we therefore
2501 find
2502 \begin{align*}
2503 D_{P(\vec p), F(\vec p)}
2505 -\sum_{n=0}^\infty
2506 \frac{b(n+1, \fractional{-n_1 v_1(\vec p) - n_2 v_2(\vec p)})}{(n+1)!}
2507 \left(\frac{n_1}d D_1 + \frac{n_2}d D_2\right)^n
2510 - \sum_{i=0}^\infty \sum_{j=0}^\infty
2511 \frac{b(i+j+1, \fractional{-n_1 v_1(\vec p) - n_2 v_2(\vec p)})}{(i+j+1)!}
2512 \frac{n_1^i n_2^j}{d^{i+j}} D_1^i D_2^j
2514 \end{align*}
2515 After applying this differential operator to the polynomial
2516 $h(\vec p)(\vec x)$, the resulting polynomial
2518 h'(\vec p)(\vec x) = D_{P(\vec p), F(\vec p)} \cdot h(\vec p)(\vec x)
2520 needs to be integrated over the facet.
2521 The measure to be used is such that the integral of a lattice tile
2522 in the linear space parallel to the facet is 1, i.e.,
2524 \int_{\vec 0}^{\vec f} 1 = \int_0^1 1 dz = 1,
2526 with $z$ the coordinate along $\vec f$.
2527 Referring to \eqref{eq:EML:facet:coordinates} and
2528 \eqref{eq:EML:facet:coordinates:2}, all points of the facet
2529 have the form $\vec x(\vec p) = z \, \vec f + a_2(\vec p) \, \vec n$,
2530 while the $z$-coordinate of the vertices $\vec v_1(\vec p)$
2531 and $\vec v_2(\vec p)$ are
2532 $(-n_2 v_{1,1} + n_1 v_{1,2})/d$
2534 $(-n_2 v_{2,1} + n_1 v_{2,2})/d$, respectively.
2535 That is, the contribution of the facet is equal to
2537 \int_{(-n_2 v_{1,1} + n_1 v_{1,2})/d}^{(-n_2 v_{2,1} + n_1 v_{2,2})/d}
2538 h'(\vec p)\left(z \, \vec f + a_2(\vec p) \, \vec n\right) \, dz
2541 where, again, we need to ensure that the lower limit is smaller
2542 than the upper limit using the usual method of plugging in a
2543 particular value of the parameters.
2545 Finally, we consider the contributions of the vertices.
2546 The \ai{transverse cone}s are in this case simply the supporting cones.
2547 Since $\mu$ is a valuation, we may apply \ai{Barvinok's decomposition}
2548 and assume that the cone is unimodular.
2549 For an affine cone
2550 \begin{align*}
2551 K &= \vec v(\vec p) + \RR_+ \vec r_1 + \RR_+ \vec r_2
2553 &= (a_1(\vec p) + \RR_+) \vec r_1 + (a_2(\vec p) + \RR_+) \vec r_2,
2554 \end{align*}
2555 with
2557 \vec a(\vec p) =
2558 \begin{bmatrix}
2559 \vec r_1 & \vec r_2
2560 \end{bmatrix}^{-1}
2561 \vec v(\vec p)
2564 we have~\cite[Proposition~31]{Berline2006local},
2565 \begin{equation}
2566 \label{eq:mu:2}
2567 \mu(K)(\vec\xi)
2569 \frac{e^{t_1 y_1 + t_2 y_2}}{(1-e^{y_1})(1-e^{y_2})}
2570 + \frac 1{y_1}B(y_2 - C_1 y_1, t_2)
2571 + \frac 1{y_2}B(y_1 - C_2 y_2, t_1)
2572 - \frac 1{y_1 y_2},
2573 \end{equation}
2574 with
2576 B(y,t) =
2577 \frac{e^{t y}}{1-e^y} + \frac 1{y}
2578 = -\sum_{n=0}^\infty \frac{b(n+1, t)}{(n+1)!} y^n
2581 $y_i = \sps{\vec\xi}{\vec r_i}$,
2582 $C_i = \sps{\vec v_1}{\vec v_2}/\sps{\vec v_i}{\vec v_i}$
2584 $t_i = \fractional{-a_i(\vec p)}$.
2585 Expanding \eqref{eq:mu:2}, we find
2586 \begin{align*}
2587 \mu(K)(\vec\xi)
2589 \left(
2590 -\frac{b(0,t1)}{y_1} - \sum_{n=0}^\infty \frac{b(n+1,t_1)}{(n+1)!} y_1^n
2591 \right)
2592 \left(
2593 -\frac{b(0,t2)}{y_2} - \sum_{n=0}^\infty \frac{b(n+1,t_2)}{(n+1)!} y_2^n
2594 \right)
2596 & \phantom{=}
2598 \left(
2599 \sum_{n=0}^\infty \frac{b(n+1,t_2)}{(n+1)!} \frac{y_2^n}{y_1}
2601 \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}
2602 \right)
2604 & \phantom{=}
2606 \left(
2607 \sum_{n=0}^\infty \frac{b(n+1,t_1)}{(n+1)!} \frac{y_1^n}{y_2}
2609 \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}
2610 \right)
2612 & \phantom{=}
2613 - \frac 1{y_1 y_2}
2616 \sum_{n_1=0}^\infty
2617 \sum_{n_2=0}^\infty
2618 c(C_1, C_2, t_1, t_2; n_1, n_2) \, y_1^n y_2^n
2620 \end{align*}
2621 with
2622 \begin{align*}
2623 c(C_1, C_2, t_1, t_2; n_1, n_2)
2625 \frac{b(n_1+1,t_1)}{(n_1+1)!} \frac{b(n_2+1,t_2)}{(n_2+1)!}
2629 \frac{b(n_1+n_2+1,t_2)}{(n_1+n_2+1)!} {n_1+n_2+1 \choose n_1+1}
2630 \left(-C_1\right)^{n_1+1}
2634 \frac{b(n_1+n_2+1,t_1)}{(n_1+n_2+1)!} {n_1+n_2+1 \choose n_2+1}
2635 \left(-C_2\right)^{n_2+1}
2637 \end{align*}
2638 For $\vec \xi = \vec D$, we have
2639 \begin{align*}
2640 y_1^n y_2^n
2642 \left( r_{1,1} D_1 + r_{1,2} D_2 \right)^{n_1}
2643 \left( r_{2,1} D_1 + r_{2,2} D_2 \right)^{n_2}
2646 \left(
2647 \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}
2648 \right)
2649 \left(
2650 \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}
2651 \right)
2652 \end{align*}
2653 and so
2655 D_{P(\vec p), \vec v(\vec p)} = \mu(K)(\vec D)
2659 \sum_{i=0}^\infty
2660 \sum_{j=0}^\infty
2661 \sum_{\shortstack{$\scriptstyle i+j = n_1+n_2$\\$\scriptstyle n_1 \ge 0$\\$\scriptstyle n_2 \ge 0$}}
2662 \sum_{\shortstack{$\scriptstyle k+l = i$\\$\scriptstyle 0 \le k \le n_1$\\$\scriptstyle 0 \le l \le n_2$}}
2663 c(C_1, C_2, t_1, t_2; n_1, n_2)
2664 r_{1,1}^k r_{1,2}^{n_1 - k}
2665 r_{2,1}^l r_{2,2}^{n_2 - l}
2666 { n_1 \choose k} { n_2 \choose l} D_1^i D_2^j
2669 The contribution of this vertex is then
2671 h'(\vec p)(\vec v(\vec p))
2674 with $
2675 h'(\vec p)(\vec x) = D_{P(\vec p), \vec v(\vec p)} \cdot h(\vec p)(\vec x)
2678 \begin{example}
2679 As a simple example, consider the (non-parametric) triangle
2680 in \autoref{f:EML:triangle} and assume we want to compute
2682 \sum_{\vec x \in T \cap \ZZ^2} x_1 x_2
2685 Since $T \cap \ZZ^2 = \{\, (2,4), (3,4), (2,5) \, \}$,
2686 the result should be
2688 2 \cdot 4 + 3 \cdot 4 + 2 \cdot 5 = 30
2692 \begin{figure}
2693 \intercol=1.2cm
2694 \begin{xy}
2695 <\intercol,0pt>:<0pt,\intercol>::
2696 \POS@i@={(2,4),(3,4),(2,5),(2,4)},{0*[|(2)]\xypolyline{}}
2697 \POS(2.35,4.25)*{x_1 x_2}
2698 \POS(2,4)*+!U{(2,4)}
2699 \POS(3,4)*+!U{(3,4)}
2700 \POS(2,5)*+!D{(2,5)}
2701 \POS(2,4)*{\cdot}
2702 \POS(3,4)*{\cdot}
2703 \POS(2,5)*{\cdot}
2704 \POS(-1,0)\ar(4,0)
2705 \POS(0,-1)\ar(0,5.5)
2706 \end{xy}
2707 \caption{Sum of polynomial $x_1 x_2$ over the integer points in a triangle $T$}
2708 \label{f:EML:triangle}
2709 \end{figure}
2711 Let us first consider the integral
2713 \iint_T x_1 x_2 = \int_{\partial T} \frac{x_1^2 x_2}2 \, d x_2
2716 Integration along each of the edges of the triangle yields
2717 the following.
2719 \marginpar{%
2720 \intercol=1cm
2721 \begin{xy}
2722 <\intercol,0pt>:<0pt,\intercol>::
2723 \POS(0,-1)*\xybox{
2724 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2725 \POS(0,0)\ar@[|(2)](1,0)
2727 \end{xy}
2729 For the edge in the margin, we have $\vec f = (1,0)$, i.e., $f_2 = 0$.
2730 The contribution of this edge to the integral is therefore zero.
2732 \marginpar{%
2733 \intercol=1cm
2734 \begin{xy}
2735 <\intercol,0pt>:<0pt,\intercol>::
2736 \POS(0,-1)*\xybox{
2737 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2738 \POS(1,0)\ar@[|(2)](0,1)
2740 \end{xy}
2742 For this edge, we have $\vec f = (-1,1)$.
2743 The contribution of this edge to the integral is therefore
2745 \int_0^1 \frac{(3-\lambda)^2(4+\lambda)}2 d\lambda
2746 = \frac{337}{24}
2750 \marginpar{%
2751 \intercol=1cm
2752 \begin{xy}
2753 <\intercol,0pt>:<0pt,\intercol>::
2754 \POS(0,-1)*\xybox{
2755 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2756 \POS(0,1)\ar@[|(2)](0,0)
2758 \end{xy}
2760 For this edge, we have $\vec f = (0,-1)$.
2761 The contribution of this edge to the integral is therefore
2763 \int_0^1 \frac{2^2(5-\lambda)}2 (-1) d\lambda
2764 = -9
2768 The total integral is therefore
2770 \int_{\partial T} \frac{x_1^2 x_2}2 \, d x_2
2771 = 0 + \frac{337}{24} - 9 = \frac{121}{24}
2775 Now let us consider the contributions of the edges.
2776 We will need the following \ai{Bernoulli number}s in our
2777 computations.
2778 \begin{align*}
2779 b(1,0) & = - \frac 1 2
2781 b(2,0) & = \frac 1 6
2783 b(3,0) & = 0
2785 b(4,0) & = -\frac 1 {30}
2786 \end{align*}
2788 \marginpar{%
2789 \intercol=1cm
2790 \begin{xy}
2791 <\intercol,0pt>:<0pt,\intercol>::
2792 \POS(0,-1)*\xybox{
2793 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2794 \POS(0,0)\ar@[|(2)]@{-}(1,0)
2795 \POS(0.5,0)\ar(0.5,1)
2797 \end{xy}
2799 The normal to the facet $F_1$ in the margin is $\vec n = (0,1)$.
2800 The vector $\vec f = (-1,0)$ is parallel to the facet.
2801 We have
2803 \begin{bmatrix}
2804 2 \\ 4
2805 \end{bmatrix}
2808 \begin{bmatrix}
2809 -1 \\ 0
2810 \end{bmatrix}
2812 \begin{bmatrix}
2813 0 \\ 1
2814 \end{bmatrix}
2815 \quad\text{and}\quad
2816 \begin{bmatrix}
2817 3 \\ 4
2818 \end{bmatrix}
2821 \begin{bmatrix}
2822 -1 \\ 0
2823 \end{bmatrix}
2825 \begin{bmatrix}
2826 0 \\ 1
2827 \end{bmatrix}
2830 Therefore $t = \fractional{-4} = 0$, $y = D_2$,
2831 \begin{align*}
2832 D_{T,F_1}
2833 & =
2834 - \sum_{j=0}^\infty \frac{b(j+1, 0)}{(j+1)!} D_2^j
2837 - \frac{b(1,0)}1 - \frac{b(2,0)}2 D_2 + \cdots
2838 \end{align*}
2841 h'(\vec x) =
2842 D_{T,F_1} \cdot x_1 x_2 =
2843 \left(\frac 1 2 - \frac 1{12} D_2\right) \cdot x_1 x_2
2845 \frac 1 2 x_1 x_2 - \frac 1{12} x_1
2848 With $x_1 = - z$ and $x_2 = 4$, the contribution of this facet
2851 \int_{-3}^{-2} - 2 z + \frac 1{12} z \, dz
2853 \frac{115}{24}
2857 \marginpar{%
2858 \intercol=1cm
2859 \begin{xy}
2860 <\intercol,0pt>:<0pt,\intercol>::
2861 \POS(0,-1)*\xybox{
2862 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2863 \POS(0,0)\ar@[|(2)]@{-}(0,1)
2864 \POS(0,0.5)\ar(1,0.5)
2866 \end{xy}
2868 The normal to the facet $F_2$ in the margin is $\vec n = (1,0)$.
2869 The vector $\vec f = (0,1)$ is parallel to the facet.
2870 We have
2872 \begin{bmatrix}
2873 2 \\ 4
2874 \end{bmatrix}
2877 \begin{bmatrix}
2878 0 \\ 1
2879 \end{bmatrix}
2881 \begin{bmatrix}
2882 1 \\ 0
2883 \end{bmatrix}
2884 \quad\text{and}\quad
2885 \begin{bmatrix}
2886 2 \\ 5
2887 \end{bmatrix}
2890 \begin{bmatrix}
2891 0 \\ 1
2892 \end{bmatrix}
2894 \begin{bmatrix}
2895 1 \\ 0
2896 \end{bmatrix}
2899 Therefore $t = \fractional{-2} = 0$, $y = D_1$,
2900 \begin{align*}
2901 D_{T,F_2}
2902 & =
2903 - \sum_{i=0}^\infty \frac{b(i+1, 0)}{(i+1)!} D_1^i
2906 - \frac{b(1,0)}1 - \frac{b(2,0)}2 D_1 + \cdots
2907 \end{align*}
2910 h'(\vec x) =
2911 D_{T,F_2} \cdot x_1 x_2 =
2912 \left(\frac 1 2 - \frac 1{12} D_1\right) \cdot x_1 x_2
2914 \frac 1 2 x_1 x_2 - \frac 1{12} x_2
2917 With $x_1 = 2$ and $x_2 = z$, the contribution of this facet
2920 \int_{4}^{5} z - \frac 1{12} z \, dz
2922 \frac{33}{8}
2926 \marginpar{%
2927 \intercol=1cm
2928 \begin{xy}
2929 <\intercol,0pt>:<0pt,\intercol>::
2930 \POS(0,-1)*\xybox{
2931 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
2932 \POS(1,0)\ar@[|(2)]@{-}(0,1)
2933 \POS(0.5,0.5)\ar(-0.5,-0.5)
2935 \end{xy}
2937 The normal to the facet $F_3$ in the margin is $\vec n = (-1,-1)$.
2938 The vector $\vec f = (1,-1)$ is parallel to the facet.
2939 We have
2941 \begin{bmatrix}
2942 3 \\ 4
2943 \end{bmatrix}
2945 -\frac 1 2
2946 \begin{bmatrix}
2947 1 \\ -1
2948 \end{bmatrix}
2949 -\frac 7 2
2950 \begin{bmatrix}
2951 -1 \\ -1
2952 \end{bmatrix}
2953 \quad\text{and}\quad
2954 \begin{bmatrix}
2955 2 \\ 5
2956 \end{bmatrix}
2958 -\frac 3 2
2959 \begin{bmatrix}
2960 1 \\ -1
2961 \end{bmatrix}
2962 -\frac 7 2
2963 \begin{bmatrix}
2964 -1 \\ -1
2965 \end{bmatrix}
2968 Therefore $t = \fractional{7} = 0$, $y = -\frac 1 2 D_1 -\frac 1 2 D_2$,
2969 \begin{align*}
2970 D_{T,F_3}
2971 & =
2972 - \sum_{i=0}^\infty \sum_{j=0}^\infty
2973 \frac{b(i+j+1, 0)}{(i+j+1)!}
2974 \frac{(-1)^{i+j}}{2^{i+j}} D_1^i D_2^j
2977 - \frac{b(1,0)}1
2978 + \frac 1 2 \frac{b(2,0)}2 D_1
2979 + \frac 1 2 \frac{b(2,0)}2 D_2 + \cdots
2980 \end{align*}
2983 h'(\vec x) =
2984 D_{T,F_4} \cdot x_1 x_2 =
2985 \left(\frac 1 2 + \frac 1{24} D_1 + \frac 1{24} D_2\right) \cdot x_1 x_2
2987 \frac 1 2 x_1 x_2 + \frac 1{24} x_2 + \frac 1{24} x_1
2990 With $x_1 = z + \frac 7 2$ and $x_2 = -z + \frac 7 2$,
2991 the contribution of this facet
2994 \int_{-\frac 3 2}^{-\frac 1 2}
2995 \frac 1 2 (z + \frac 7 2)(-z + \frac 7 2)
2996 + \frac 1{24}(-z + \frac 7 2)
2997 + \frac 1{24}(z + \frac 7 2) \, dz
2999 \frac{47}{8}
3003 The total contribution of the edges is therefore
3005 \frac{115}{24}+\frac{33}8+
3006 \frac{47}{8} = \frac{355}{24}
3010 Finally, we consider the contributions of the vertices.
3012 \marginpar{%
3013 \intercol=1cm
3014 \begin{xy}
3015 <\intercol,0pt>:<0pt,\intercol>::
3016 \POS(0,-1)*\xybox{
3017 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
3018 \POS(1,0)\ar@[|(2)](0,1)
3019 \POS(1,0)\ar@[|(2)](0,0)
3021 \end{xy}
3023 For the vertex $\vec v = (3,4)$, we have
3024 $\vec r_1 = (-1,0)$ and $\vec r_2 = (-1,1)$.
3025 Since $\vec v$ is integer, we have $t_1 = t_2 = 0$.
3026 Also, $C_1 = 1$, $C_2 = 1/2$, $y_1 = -D_1$ and $y_2 = -D_1 + D_2$.
3027 Since the total degree of the polynomial $x_1 x_2$ is two,
3028 we only need the coefficients of $\mu(K)(\vec \xi)$ up to
3029 $n_1+n_2 = 2$.
3031 \noindent
3032 \begin{tabular}{c|c|c}
3033 $n_1$ & $n_2$
3035 \hline
3036 0 & 0 &
3038 \left(
3039 \frac{b(1,0)}{1!}
3040 \frac{b(1,0)}{1!}
3042 \frac{b(2,0)}{2!}
3043 {1 \choose 1}(-1)^1
3045 \frac{b(2,0)}{2!}
3046 {1 \choose 1}(-\frac 12)^1
3047 \right)
3050 1 & 0 &
3052 \left(
3053 \frac{b(2,0)}{2!}
3054 \frac{b(1,0)}{1!}
3056 \frac{b(3,0)}{3!}
3057 {2 \choose 2}(-1)^2
3059 \frac{b(3,0)}{3!}
3060 {2 \choose 1}(-\frac 12)^1
3061 \right)
3062 \left(
3063 -D_1
3064 \right)
3067 0 & 1 &
3069 \left(
3070 \frac{b(1,0)}{1!}
3071 \frac{b(2,0)}{2!}
3073 \frac{b(3,0)}{3!}
3074 {2 \choose 1}(-1)^1
3076 \frac{b(3,0)}{3!}
3077 {2 \choose 2}(-\frac 12)^2
3078 \right)
3079 \left(
3080 -D_1 + D_2
3081 \right)
3084 2 & 0 &
3086 \left(
3087 \frac{b(3,0)}{3!}
3088 \frac{b(1,0)}{1!}
3090 \frac{b(4,0)}{4!}
3091 {3 \choose 3}(-1)^3
3093 \frac{b(4,0)}{4!}
3094 {3 \choose 1}(-\frac 12)^1
3095 \right)
3096 \left(
3097 -D_1
3098 \right)^2
3101 1 & 1 &
3103 \left(
3104 \frac{b(2,0)}{2!}
3105 \frac{b(2,0)}{2!}
3107 \frac{b(4,0)}{4!}
3108 {3 \choose 2}(-1)^2
3110 \frac{b(4,0)}{4!}
3111 {3 \choose 2}(-\frac 12)^2
3112 \right)
3113 \left(
3114 -D_1
3115 \right)
3116 \left(
3117 -D_1 + D_2
3118 \right)
3121 0 & 2 &
3123 \left(
3124 \frac{b(1,0)}{1!}
3125 \frac{b(3,0)}{3!}
3127 \frac{b(4,0)}{4!}
3128 {3 \choose 1}(-1)^1
3130 \frac{b(4,0)}{4!}
3131 {3 \choose 3}(-\frac 12)^3
3132 \right)
3133 \left(
3134 -D_1 + D_2
3135 \right)^2
3137 \end{tabular}
3139 We find
3140 \begin{align*}
3141 h'(\vec x)
3143 \left(
3144 \frac 3 8 - \frac 1{24} (-D_1) - \frac 1{24} (-D_1 + D_2)
3145 + \frac 7{576} (-D_1 D_2)
3146 - \frac 5{1152} (-2 D_1 D2)
3147 \right) x_1 x_2
3150 \frac 3 8 x_1 x_2 + \frac 1{24} x_2 - \frac 1{24} (-x_2 + x_1)
3151 + \frac 7{576} (-1)
3152 - \frac 5{1152} (-2)
3154 \end{align*}
3155 The contribution of this vertex is therefore
3157 h'(3,4) = \frac {1355}{288}
3161 \marginpar{%
3162 \intercol=1cm
3163 \begin{xy}
3164 <\intercol,0pt>:<0pt,\intercol>::
3165 \POS(0,-1)*\xybox{
3166 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
3167 \POS(0,1)\ar@[|(2)](1,0)
3168 \POS(0,1)\ar@[|(2)](0,0)
3170 \end{xy}
3172 For the vertex $\vec v = (2,5)$, we have
3173 $\vec r_1 = (0,-1)$ and $\vec r_2 = (1,-1)$.
3174 Since $\vec v$ is integer, we have $t_1 = t_2 = 0$.
3175 Also, $C_1 = 1$, $C_2 = 1/2$, $y_1 = -D_2$ and $y_2 = D_1 - D_2$.
3176 We similarly find
3178 h'(\vec x)
3180 \frac 3 8 x_1 x_2 + \frac 1{24} x_1 - \frac 1{24} (x_2 - x_1)
3181 + \frac 7{576} (-1)
3182 - \frac 5{1152} (-2)
3185 The contribution of this vertex is therefore
3187 h'(2,5) = \frac {1067}{288}
3191 \marginpar{%
3192 \intercol=1cm
3193 \begin{xy}
3194 <\intercol,0pt>:<0pt,\intercol>::
3195 \POS(0,-1)*\xybox{
3196 \POS@i@={(0,0),(1,0),(0,1),(0,0)},{0*\xypolyline{--}}
3197 \POS(0,0)\ar@[|(2)](1,0)
3198 \POS(0,0)\ar@[|(2)](0,1)
3200 \end{xy}
3202 For the vertex $\vec v = (2,4)$, we have
3203 $\vec r_1 = (1,0)$ and $\vec r_2 = (0,1)$.
3204 Since $\vec v$ is integer, we have $t_1 = t_2 = 0$.
3205 The computations are easier in this case since
3206 $C_1 = C_2 = 0$, $y_1 = D_1$ and $y_2 = D_2$.
3207 We find
3209 h'(\vec x)
3211 \frac 1 4 x_1 x_2 - \frac 1{12} x_2 - \frac 1{12} x_1
3212 + \frac 1{144} (1)
3215 The contribution of this vertex is therefore
3217 h'(2,4) = \frac {253}{144}
3221 The total contribution of the vertices is then
3223 \frac {1355}{288} + \frac {1067}{288} + \frac {253}{144}
3224 = \frac {61}6
3226 and the total sum is
3228 \frac{121}{24}+\frac{355}{24}+\frac{61}6 = 30
3232 \end{example}
3234 \begin{example}
3235 Consider the parametric polytope
3237 P(n) = \{\, \vec x \mid x_1 \ge 2 \wedge 3 x_1 \le n + 9
3238 \wedge 4 \le x_2 \le 5 \,\}
3241 If $n \ge -3$, then the vertices of this polytope are
3242 $(2,4)$, $(2,5)$, $(3+n/3,4)$ and $(3+n/3,5)$.
3243 The contributions of the faces of $P(n)$ to
3245 \sum_{\vec x \in P(n) \cap \ZZ^2} x_1 x_2
3247 for the chamber $n \ge -3$ are shown in \autoref{t:sum:rectangle}.
3248 The final result is
3250 \begin{cases}
3251 \frac{ n^2}{2}
3252 - 3 n \fractional{\frac{ n}{3}}
3253 + \frac{21}{2} n
3254 + \frac{9}{2} \fractional{\frac{ n}{3}}^2
3255 - \frac{63}{2} \fractional{\frac{ n}{3}}
3256 + 45
3257 & \text{if $ n+3 \ge 0$}.
3258 \end{cases}
3261 \begin{table}
3262 \intercol=1cm
3263 \begin{tabular}{lc}
3264 \begin{xy}
3265 <\intercol,0pt>:<0pt,\intercol>::
3266 \POS(-1,-0.5)*\xybox{
3267 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*[|(2)]\xypolyline{}}
3268 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3270 \end{xy}
3273 \displaystyle
3274 \frac{ n^2}{4}
3275 + \frac{9}{2} n
3276 + \frac{45}{4}
3279 \begin{xy}
3280 <\intercol,0pt>:<0pt,\intercol>::
3281 \POS(-1,-0.5)*\xybox{
3282 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3283 \POS(0,0)\ar@[|(2)]@{-}(0,1)
3284 \POS(0,0.5)*+!L{2}
3285 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3287 \end{xy}
3290 \displaystyle
3291 \frac{33}{8}
3294 \begin{xy}
3295 <\intercol,0pt>:<0pt,\intercol>::
3296 \POS(-1,-0.5)*\xybox{
3297 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3298 \POS(1,0)\ar@[|(2)]@{-}(1,1)
3299 \POS(1,0.5)*+!L{3+n/3}
3300 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3302 \end{xy}
3305 \displaystyle
3306 - \frac{3}{2} n \fractional{\frac{ n}{3}}
3307 + \frac{3}{4} n
3308 + \frac{9}{4} \fractional{\frac{ n}{3}}^2
3309 - \frac{63}{4} \fractional{\frac{ n}{3}}
3310 + \frac{57}{8}
3313 \begin{xy}
3314 <\intercol,0pt>:<0pt,\intercol>::
3315 \POS(-1,-0.5)*\xybox{
3316 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3317 \POS(0,0)\ar@[|(2)]@{-}(1,0)
3318 \POS(0.5,0)*+!D{4}
3319 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3321 \end{xy}
3324 \displaystyle
3325 \frac{23}{216} n^2
3326 + \frac{23}{12} n
3327 + \frac{115}{24}
3330 \begin{xy}
3331 <\intercol,0pt>:<0pt,\intercol>::
3332 \POS(-1,-0.5)*\xybox{
3333 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3334 \POS(0,1)\ar@[|(2)]@{-}(1,1)
3335 \POS(0.5,1)*+!U{5}
3336 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3338 \end{xy}
3341 \displaystyle
3342 \frac{31}{216} n^2
3343 + \frac{31}{12} n
3344 + \frac{155}{24}
3347 \begin{xy}
3348 <\intercol,0pt>:<0pt,\intercol>::
3349 \POS(-1,-0.5)*\xybox{
3350 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3351 \POS(1,1)\ar@[|(2)](1,0)
3352 \POS(1,1)\ar@[|(2)](0,1)
3353 \POS(1,1)*+!LU{(3+n/3,5)}
3354 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3356 \end{xy}
3359 \displaystyle
3360 - \frac{31}{36} n \fractional{\frac{ n}{3}}
3361 + \frac{31}{72} n
3362 + \frac{31}{24} \fractional{\frac{ n}{3}}^2
3363 - \frac{217}{24} \fractional{\frac{ n}{3}}
3364 + \frac{589}{144}
3367 \begin{xy}
3368 <\intercol,0pt>:<0pt,\intercol>::
3369 \POS(-1,-0.5)*\xybox{
3370 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3371 \POS(0,1)\ar@[|(2)](1,1)
3372 \POS(0,1)\ar@[|(2)](0,0)
3373 \POS(0,1)*+!LU{(2,5)}
3374 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3376 \end{xy}
3379 \displaystyle
3380 \frac{341}{144}
3383 \begin{xy}
3384 <\intercol,0pt>:<0pt,\intercol>::
3385 \POS(-1,-0.5)*\xybox{
3386 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3387 \POS(0,0)\ar@[|(2)](1,0)
3388 \POS(0,0)\ar@[|(2)](0,1)
3389 \POS(0,0)*+!LD{(2,4)}
3390 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3392 \end{xy}
3395 \displaystyle
3396 \frac{253}{144}
3399 \begin{xy}
3400 <\intercol,0pt>:<0pt,\intercol>::
3401 \POS(-1,-0.5)*\xybox{
3402 \POS@i@={(0,0),(1,0),(1,1),(0,1),(0,0)},{0*\xypolyline{--}}
3403 \POS(1,0)\ar@[|(2)](1,1)
3404 \POS(1,0)\ar@[|(2)](0,0)
3405 \POS(1,0)*+!LD{(3+n/3,4)}
3406 \POS(0,1.1)*{}\POS(0,-0.1)*{}
3408 \end{xy}
3411 \displaystyle
3412 - \frac{23}{36} n \fractional{\frac{ n}{3}}
3413 + \frac{23}{72} n
3414 + \frac{23}{24} \fractional{\frac{ n}{3}}^2
3415 - \frac{161}{24} \fractional{\frac{ n}{3}}
3416 + \frac{437}{144}
3418 \end{tabular}
3419 \caption{Contributions of the faces of $P(n)$ to the sum of $x_1 x_2$ over
3420 the integer points of $P(n)$}
3421 \label{t:sum:rectangle}
3422 \end{table}
3424 \end{example}
3427 \subsection{Conversion to ``standard form''}
3428 \label{s:standard}
3430 Some algorithms or tools expect a polyhedron to be
3431 specified in ``\ai{standard form}'', i.e.,
3432 \begin{equation}
3433 \label{eq:standard}
3434 \begin{cases}
3435 \begin{aligned}
3436 A \vec x & = \vec b \\
3437 \vec x & \ge \vec 0
3439 \end{aligned}
3440 \end{cases}
3441 \end{equation}
3442 Given an arbitrary (parametric) polyhedron
3443 \begin{equation}
3444 \label{eq:non-standard}
3445 \{\,
3446 \vec x \mid
3447 A \vec x + \vec b(\vec p) \ge 0
3448 \,\}
3450 \end{equation}
3451 a conversion to standard form requires the introduction
3452 of \ai{slack variable}s and a way of dealing with variables
3453 of \ai{unrestricted sign}.
3454 In this section we will be satisfied with a reduction
3455 to the form
3456 \begin{equation}
3457 \label{eq:standard:2}
3458 \begin{cases}
3459 \begin{aligned}
3460 A \vec x & = \vec b \\
3461 D \vec x & \ge \vec c
3463 \end{aligned}
3464 \end{cases}
3465 \end{equation}
3466 with $D$ a diagonal matrix with positive entries.
3467 That is, we do not necessarily make all variables non-negative,
3468 but we do ensure that they have a lower bound.
3469 If needed, a subsequent reduction can then be performed.
3471 The standard way of dealing with variables of unrestricted
3472 sign is to replace a variable $x$ of unknown sign by the
3473 difference ($x = x' - x''$) of two non-negative variables
3474 ($x', x'' \ge 0$).
3475 However, some algorithms are somewhat sensitive with respect
3476 to the number of variables and so we would prefer to introduce
3477 as few extra variables as possible.
3478 We will therefore apply a \ai{unimodular transformation}
3479 on the variables such that all transformed variables are known
3480 to be non-negative.
3482 The first step is to compute the \indac{HNF} of A,
3483 i.e., a matrix $H = A U$, with $U$ unimodular,
3484 in column echelon form such that the
3485 first entry in each column is positive and the other entries
3486 on the corresponding row are non-negative and strictly smaller
3487 than this first entry.
3488 By reordering the rows we may assume that the top square part
3489 of $H$ is lower-triangular.
3490 By a further unimodular transformation, the entries
3491 below the diagonal can be made non-positive and strictly
3492 smaller (in absolute value) than the diagonal entry of the same row.
3494 For each of the new variables, we can take a positive
3495 combination of the corresponding row and the previous rows
3496 to obtain a positive multiple of the corresponding unit vector,
3497 implying that the variable has a lower bound.
3498 A slack variable can then be introduced for each of the
3499 rows in the top square part of $H'$ that is not already
3500 a positive multiple of a unit vector and for each of
3501 the rows below the top square part of $H'$.
3503 \begin{example}
3504 Consider the cone
3506 \left\{\,
3507 \vec x \mid
3508 \begin{bmatrix}
3509 67 & -13 \\
3510 -52 & 53
3511 \end{bmatrix}
3512 \vec x
3514 \vec 0
3515 \,\right\}
3518 This cone is already situated in the first quadrant,
3519 but this may not be obvious from the constraints.
3520 Furthermore, directly adding slack variables would
3521 lead to a total of 4 variables, whereas we can also
3522 represent this cone in standard form with only 3 variables.
3523 We have
3525 H' =
3526 \begin{bmatrix}
3527 1 & 0 \\
3528 -1331 & 2875
3529 \end{bmatrix}
3531 \begin{bmatrix}
3532 67 & -13 \\
3533 -52 & 53
3534 \end{bmatrix}
3535 \begin{bmatrix}
3536 -6 & 13 \\
3537 -31 & 57
3538 \end{bmatrix}
3539 = A U'
3542 Adding a slack variable for the second row of $H'$, we
3543 obtain the equivalent problem
3545 \begin{cases}
3546 \begin{aligned}
3547 \begin{bmatrix}
3548 -1331 & 2875 & -1
3549 \end{bmatrix}
3550 \vec x'
3552 \vec 0
3554 \vec x' & \ge \vec 0
3555 \end{aligned}
3556 \end{cases}
3558 with
3560 \vec x =
3561 \begin{bmatrix}
3562 -6 & 13 & 0 \\
3563 -31 & 57 & 0
3564 \end{bmatrix}
3565 \vec x'
3568 \end{example}
3570 A similar construction was used by \shortciteN[Lemma~3.10]{Eisenbrand2000PhD}
3571 and \shortciteN{Hung1990}.
3573 \subsection{Using TOPCOM to compute Chamber Decompositions}
3575 In this section, we describe how to use the correspondence
3576 between the \ai{regular triangulation}s of a point set
3577 and the chambers of the \ai{Gale transform}
3578 of the point set~\shortcite{Gelfand1994}
3579 to compute the chamber decomposition of a parametric polytope.
3580 This correspondence was also used by \shortciteN{Pfeifle2003}
3581 \shortciteN{Eisenschmidt2007integrally}.
3583 Let us first assume that the parametric polytope can be written as
3584 \begin{equation}
3585 \label{eq:TOPCOM:polytope}
3586 \begin{cases}
3587 \begin{aligned}
3588 \vec x &\ge 0
3590 A \, \vec x &\le \vec b(\vec p)
3592 \end{aligned}
3593 \end{cases}
3594 \end{equation}
3595 where the right hand side $\vec b(\vec p)$ is arbitrary and
3596 may depend on the parameters.
3597 The first step is to add slack variables $\vec s$ to obtain
3598 the \ai{vector partition} problem
3600 \begin{cases}
3601 \begin{aligned}
3602 A \, \vec x + I \, \vec s & = \vec b(\vec p)
3604 \vec x, \vec s &\ge 0
3606 \end{aligned}
3607 \end{cases}
3609 with $I$ the identity matrix.
3610 Then we compute the (right) kernel $K$ of the matrix
3611 $\begin{bmatrix}
3612 A & I
3613 \end{bmatrix}$, i.e.,
3615 \begin{bmatrix}
3616 A & I
3617 \end{bmatrix}
3622 and use \ai[\tt]{TOPCOM}'s \ai[\tt]{points2triangs} to
3623 compute the \ai{regular triangulation}s of the points specified
3624 by the rows of $K$.
3625 Each of the resulting triangulations corresponds to a chamber
3626 in the chamber complex of the above vector partition problem.
3627 Each simplex in a triangulation corresponds to a parametric
3628 vertex active on the corresponding chamber and
3629 each point in the simplex (i.e., a row of $K$) corresponds
3630 to a variable ($x_j$ or $s_j$) that is set to zero to obtain
3631 this parametric vertex.
3632 In the original formulation of the problem~\eqref{eq:TOPCOM:polytope}
3633 each such variable set to zero reflects the saturation of the
3634 corresponding constraint ($x_j = 0$ for $x_j = 0$ and
3635 $\sps {\vec a_j}{\vec x} = b_j(\vec p)$ for $s_j = 0$).
3636 A description of the chamber can then be obtained by plugging
3637 in the parametric vertices in the remaining constraints.
3639 \begin{example}
3640 Consider the parametric polytope
3642 P(p,q,r) = \{\,
3643 (i,j) \mid 0 \le i \le p \wedge
3644 0 \le j \le 2 i + q \wedge
3645 0 \le k \le i - p + r \wedge
3646 p \ge 0 \wedge
3647 q \ge 0 \wedge
3648 r \ge 0
3649 \,\}
3652 The constraints involving the variables are
3654 \begin{cases}
3655 \begin{aligned}
3656 \begin{bmatrix}
3661 & & 1
3662 \end{bmatrix}
3663 \begin{bmatrix}
3664 i \\ j \\ k
3665 \end{bmatrix}
3667 \begin{matrix}
3673 \end{matrix}
3674 \begin{array}{l}
3680 \end{array}
3682 \begin{bmatrix}
3683 1 & 0 & 0
3685 -1 & 0 & 1
3687 -2 & 1 & 0
3688 \end{bmatrix}
3689 \begin{bmatrix}
3690 i \\ j \\ k
3691 \end{bmatrix}
3693 \begin{matrix}
3699 \end{matrix}
3700 \begin{array}{l}
3705 -p + r
3706 \end{array}
3707 \end{aligned}
3708 \end{cases}
3710 We have
3712 \begin{bmatrix}
3713 1 & 0 & 0 & 1 & 0 & 0 \\
3714 -1 & 0 & 1 & 0 & 1 & 0 \\
3715 -2 & 1 & 0 & 0 & 0 & 1 \\
3716 \end{bmatrix}
3717 \begin{bmatrix}
3718 -1 & 0 & 0 \\
3719 -2 & 0 & -1 \\
3720 -1 & -1 & 0 \\
3721 1 & 0 & 0 \\
3722 0 & 1 & 0 \\
3723 0 & 0 & 1 \\
3724 \end{bmatrix}
3728 Computing the \ai{regular triangulation}s of the rows of $K$
3729 using \ai[\tt]{TOPCOM}, we obtain
3730 \begin{verbatim}
3731 > cat e2.topcom
3733 [ -1 0 0 ]
3734 [ -2 0 -1 ]
3735 [ -1 -1 0 ]
3736 [ 1 0 0 ]
3737 [ 0 1 0 ]
3738 [ 0 0 1 ]
3740 > points2triangs --regular < e2.topcom
3741 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}};
3742 T[2]:={{1,2,3},{1,3,4},{2,3,5},{3,4,5},{1,2,5},{1,4,5}};
3743 T[3]:={{1,2,3},{1,3,4},{2,3,5},{3,4,5},{1,2,4},{2,4,5}};
3744 \end{verbatim}
3746 We see that we have three chambers in the decomposition,
3747 one with 8 vertices and two with 6 vertices.
3748 Take the second vertex (``\verb+{1,2,3}+'') of the first chamber.
3749 This vertex corresponds
3750 to the saturation of the constraints $j \ge 0$, $k \ge 0$
3751 and $i \le p$, i.e., $(i,j,k) = (p,0,0)$. Plugging in this
3752 vertex in the remaining constraints, we see that it is only valid
3753 in case $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$.
3754 For the remaining vertices of the first chamber, we similarly find
3756 \begin{tabular}{ccc}
3757 % e0
3758 \verb+{0,1,2}+ & $(0,0,0)$ & $p \ge 0$, $-q + r \ge 0$ and $q \ge 0$
3760 % 70
3761 \verb+{1,2,3}+ & $(p,0,0)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3763 % c8
3764 \verb+{0,1,4}+ & $(0,0,-p+r)$ & $-q + r \ge 0$, $p \ge 0$ and $q \ge 0$
3766 % 58
3767 \verb+{1,3,4}+ & $(p,0,r)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3769 % a4
3770 \verb+{0,2,5}+ & $(0,q,0)$ & $q \ge 0$, $p \ge 0$ and $-q + r \ge 0$
3772 % 34
3773 \verb+{2,3,5}+ & $(p, 2p+q, 0)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3775 % 8c
3776 \verb+{0,4,5}+ & $(0, q, -p+r)$ & $q \ge 0$, $-q + r \ge 0$ and $p \ge 0$
3778 % 1c
3779 \verb+{3,4,5}+ & $(p, 2p+q, r)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3780 \end{tabular}
3782 Combining these constraints with the initial constraints of the problem
3783 on the parameters
3784 $p \ge 0$, $q \ge 0$ and $r \ge 0$, we find the chamber
3786 \{\,
3787 (p,q,r) \mid p \ge 0 \wedge -p + r \ge 0 \wedge q \ge 0
3788 \,\}
3790 For the second chamber, we have
3792 \begin{tabular}{ccc}
3793 % 70
3794 \verb+{1,2,3}+ & $(p,0,0)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3796 % 58
3797 \verb+{1,3,4}+ & $(p,0,r)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3799 % 34
3800 \verb+{2,3,5}+ & $(p, 2p+q, 0)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3802 % 1c
3803 \verb+{3,4,5}+ & $(p, 2p+q, r)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3805 % 64
3806 \verb+{1,2,5}+ & $(-\frac q 2,0,0)$ &
3807 $-q \ge 0$, $2p + q \ge 0$ and $-2p -q+2r \ge 0$
3809 % 4c
3810 \verb+{1,4,5}+ & $(-\frac q 2,0,-p-\frac q 2+r)$ &
3811 $-q \ge 0$, $-2p -q+2r \ge 0$ and $2p + q \ge 0$
3812 \end{tabular}
3814 The chamber is therefore
3816 \{\,
3817 (p,q,r) \mid q = 0 \wedge p \ge 0 \wedge -p +r \ge 0
3818 \,\}
3820 Note that by intersecting with the initial constraints this chamber
3821 is no longer full-dimensional and can therefore be discarded.
3822 Finally, for the third chamber, we have
3824 \begin{tabular}{ccc}
3825 % 70
3826 \verb+{1,2,3}+ & $(p,0,0)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3828 % 58
3829 \verb+{1,3,4}+ & $(p,0,r)$ & $p \ge 0$, $r \ge 0$ and $2p + q \ge 0$
3831 % 34
3832 \verb+{2,3,5}+ & $(p, 2p+q, 0)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3834 % 1c
3835 \verb+{3,4,5}+ & $(p, 2p+q, r)$ & $p \ge 0$, $2p + q \ge 0$ and $r \ge 0$
3837 % 68
3838 \verb+{1,2,4}+ & $(p-r,0,0)$ &
3839 $p -r \ge 0$, $r \ge 0$ and $2p +q -2r \ge 0$
3841 % 2c
3842 \verb+{2,4,5}+ & $(p-r,2p+q-2r, 0)$ &
3843 $p -r \ge 0$, $2p +q -2r \ge 0$ and $r \ge 0$
3844 \end{tabular}
3846 The chamber is therefore
3848 \{\,
3849 (p,q,r) \mid p - r \ge 0 \wedge q \ge 0 \wedge r \ge 0
3850 \,\}
3852 \end{example}
3854 Now let us consider general parametric polytopes.
3855 First note that we can follow the same procedure as above
3856 if we replace $\vec x$ by $\vec x' - \vec c(\vec p)$
3857 in \eqref{eq:TOPCOM:polytope}, i.e.,
3858 if our problem has the form
3859 \begin{equation}
3860 \label{eq:TOPCOM:polytope:2}
3861 \begin{cases}
3862 \begin{aligned}
3863 \vec x' &\ge \vec c(\vec p)
3865 A \, \vec x' &\le \vec b(\vec p) + A \vec c(\vec p)
3867 \end{aligned}
3868 \end{cases}
3869 \end{equation}
3870 as saturating a constraint $x_i = 0$ is equivalent
3871 to saturating the constraint $x_i' = c_i(\vec p)$
3872 and, similarly, $\sps {\vec a_j}{\vec x} = b_j(\vec p)$
3873 is equivalent to
3874 $\sps {\vec a_j}{\vec x'} = b_j(\vec p) + \sps {\vec a_j}{\vec c(\vec p)}$.
3876 In the general case, the problem has the form
3878 A \vec x \ge \vec b(\vec p)
3880 and then we apply the technique of \autoref{s:standard}.
3881 Let $A'$ be a non-singular square submatrix of $A$ with the same number
3882 of columns and compute the (left) \indac{HNF} $H = A' U$ with $U$ unimodular
3883 and $H$ lower-triangular with non-positive elements below the diagonal.
3884 Replacing $\vec x$ by $U \vec x'$, we obtain
3886 \begin{cases}
3887 \begin{aligned}
3888 H \vec x' &\ge \vec b'(\vec p)
3890 -A''U \, \vec x' &\le -\vec b''(\vec p)
3892 \end{aligned}
3893 \end{cases}
3895 with $A''$ the remaining rows of $A$ and $\vec b(\vec p)$ split
3896 in the same way.
3897 If $H$ happens to be the identity matrix, then our problem is
3898 of the form \eqref{eq:TOPCOM:polytope:2} and we already know how
3899 to solve this problem.
3900 Note that, again, saturating any of the transformed constraints
3901 in $\vec x'$ is equivalent to saturating the corresponding constraint
3902 in $\vec x$. We therefore only need to compute $-A'' U$ for the
3903 computation of the kernel $K$. To construct the parametric vertices
3904 in the original coordinate system, we can simply use the original
3905 constraints.
3906 The same reasoning holds if $H$ is any diagonal matrix, since
3907 we can virtually replace $H \vec x$ by $\vec x'$ without affecting
3908 the non-negativity of the variables.
3910 If $H$ is not diagonal, then we can introduce new constraints
3911 $x'_j \ge d(\vec p)$, where $d(\vec p)$ is some symbolic constant.
3912 These constraints do not remove any solutions
3913 since each row in $H$ expresses that the corresponding variable is
3914 greater than or equal to a non-negative combination of the
3915 previous variables plus some constant.
3916 We can then proceed as before. However, to reduce unnecessary computations
3917 we may remove from $K$ the rows that correspond to these new rows.
3918 Any solution saturating the new constraint, would also saturate
3919 the corresponding constraint $\vec h_j^\T$ and all
3920 the constraints corresponding to the non-zero
3921 entries in $\vec h_j^\T$.
3922 If a chamber contains a vertex obtained by saturating such a new
3923 constraint, it would appear multiple times in the same chamber,
3924 each time combined with different constraints from the above set.
3925 Furthermore, there would also be another (as it turns out, identical)
3926 chamber where the vertex is only defined by the other constraints.
3928 \begin{example}
3929 Consider the parametric polytope
3931 P(n) = \{\,
3932 (i,j) \mid
3933 1 \le i \wedge 2 i \le 3 j \wedge j \le n
3934 \,\}
3937 The constraints are
3939 \begin{bmatrix}
3940 1 & 0 \\
3941 -2 & 3 \\
3942 0 & -1
3943 \end{bmatrix}
3944 \begin{bmatrix}
3945 i \\ j
3946 \end{bmatrix}
3948 \begin{bmatrix}
3949 1 \\
3950 0 \\
3952 \end{bmatrix}
3955 The top $2 \times 2$ submatrix is already in \indac{HNF}.
3956 We have $3 j \ge 2i \ge 2$, so we can add a constraint
3957 of the form $j \ge c(n)$ and obtain
3959 \begin{bmatrix}
3960 A & I
3961 \end{bmatrix}
3963 \begin{bmatrix}
3964 0 & 1 & 1 & 0
3966 2 & -3 & 0 & 1
3967 \end{bmatrix}
3970 while $K$ with $\begin{bmatrix}A & I\end{bmatrix} K = 0$ is given
3973 \begin{bmatrix}
3974 0 & 1 & 1 & 0
3976 2 & -3 & 0 & 1
3977 \end{bmatrix}
3978 \begin{bmatrix}
3979 1 & 0 \\
3980 0 & 1 \\
3981 0 & -1 \\
3982 -2 & 3
3983 \end{bmatrix}
3986 The second row of $K$ corresponds to the second variable,
3987 which in turn corresponds to the newly added constraint.
3988 Passing all rows of $K$ to \ai[\tt]{TOPCOM} we would get
3989 \begin{verbatim}
3990 > points2triangs --regular <<EOF
3991 > [[1 0],[0,1],[0,-1],[-2,3]]
3992 > EOF
3993 T[1]:={{0,1},{0,2},{1,3},{2,3}};
3994 T[2]:={{0,2},{2,3},{0,3}};
3995 T[3]:={};
3996 \end{verbatim}
3997 The first vertex in the first chamber saturates the second row
3998 (row 1) and therefore saturates both the first (0) and fourth (3)
3999 and it appears a second time as \verb+{1,3}+. Combining
4000 these ``two'' vertices into one as \verb+{0,3}+ results in the
4001 second (identical) chamber.
4002 Removing the row corresponding to the new constraint from $K$
4003 we remove the duplicates
4004 \begin{verbatim}
4005 > points2triangs --regular <<EOF
4006 > [[1 0],[0,-1],[-2,3]]
4007 > EOF
4008 T[1]:={{0,1},{1,2},{0,2}};
4009 T[2]:={};
4010 \end{verbatim}
4011 Note that in this example, we also could have interchanged
4012 the second and the third constraint and then have replaced $j$ by $-j'$.
4013 \end{example}
4015 In practice, this method of computing a \ai{chamber decomposition}
4016 does not seem to perform very well, mostly because
4017 \ai[\tt]{TOPCOM} can not exploit all available information
4018 about the parametric polytopes and will therefore compute
4019 many redundant triangulations/chambers.
4020 In particular, any chamber that does not intersect with
4021 the parameter domain of the parametric polytope, or only
4022 intersects in a face of this parameter domain, is completely redundant.
4023 Furthermore, if the parametric polytope is not simple, then many
4024 different combinations of the constraints will lead to the same parametric
4025 vertex. Many triangulations will therefore correspond to one and the
4026 same chamber in the chamber complex of the parametric polytope.
4027 For example, for a dilated octahedron, \ai[\tt]{TOPCOM} will
4028 compute 150 triangulations/chambers, 104 of which are empty,
4029 while the remaining 46 refer to the same single chamber.
4032 \subsection{Computing the Hilbert basis of a cone}
4033 \label{s:hilbert}
4035 To compute the \ai{Hilbert basis} of a cone, we use
4036 the \ai[\tt]{zsolve} library from \ai[\tt]{4ti2} \shortcite{4ti2},
4037 which implements the technique of \shortciteN{Hemmecke2002Hilbert}.
4038 We first remove all equalities from the cone through unimodular
4039 transformations and then apply the technique of \autoref{s:standard}
4040 to put the cone in ``standard form''. Note that for a (non-parametric)
4041 cone the constant term $\vec b$ in \eqref{eq:non-standard} is $\vec 0$.
4042 The constraints $D \vec x \ge \vec c = \vec 0$ of \eqref{eq:standard:2}
4043 are therefore equivalent to $\vec x \ge \vec 0$.
4046 \subsection{Integer Feasibility}
4047 \label{s:feasibility}
4049 For testing whether a polytope $P \subset \QQ^d$ contains any integer points,
4050 we use the technique of~\shortciteN{Cook1993implementation},
4051 based on \ai{generalized basis reduction}.
4053 The technique basically looks for a ``short vector'' $\vec c$ in the
4054 lattice $\ZZ^d$, where shortness is measured in terms of
4055 the \ai{width} of the polytope $P$ along that direction,
4056 \begin{align*}
4057 \width_{\vec c} P
4059 \max \{\, \sp c x \mid \vec x \in P \,\}
4061 \min \{\, \sp c x \mid \vec x \in P \,\}
4064 \max \{\, \sps {\vec c} {\vec x - \vec y} \mid \vec x, \vec y \in P \,\}
4066 \end{align*}
4067 The \defindex{lattice width} is the minimum width over all
4068 non-zero integer directions:
4070 \width P =
4071 \min_{\vec c \in \ZZ^d \setminus \{ \vec 0 \} } \width_{\vec c} P
4074 If the dimension $d$ is fixed then
4075 the lattice width of any polytope $P \subset \QQ^d$
4076 containing no integer points is bounded by a constant%
4077 ~\shortcite{Lagarias90,Barvinok02,Banaszczyk1999flatness}.
4078 If we slice the polytope using hyperplanes orthogonal
4079 to a short direction, i.e., a direction where the width
4080 is small, we will therefore only need to inspect
4081 ``few'' of them before either finding one with an integer point,
4082 or running out of hyperplanes, meaning that the
4083 polytope did not contain any integer points.
4084 Each slice is checked for integer points by applying
4085 the above method recursively.
4087 A nice feature of this technique is that it will
4088 not only tell you if there is any integer point
4089 in the given polytope, but it will actually compute
4090 one if there is any.
4092 The short vector is obtained as the first vector
4093 of a ``reduced basis'' of the lattice $\ZZ^d$ with respect
4094 to the polytope.
4095 In particular, the first vector $\vec b_1$ of this reduced basis
4096 will satisfy
4098 \width_{\vec b_1} P
4100 \frac{\width P}
4101 {\left(\frac 1 2 - \varepsilon\right)^{d-1}}
4104 with $0 < \varepsilon < 1/2$ a fixed constant.
4105 That is, the width in direction $\vec b_1$ is no more than a constant
4106 factor bigger than the lattice width.
4107 See~\shortcite{Cook1993implementation} for details.
4108 In our implementation we use $\varepsilon = 1/4$.
4109 When used in the above integer feasibility testing algorithm,
4110 we will also terminate the reduced basis computation
4111 as soon as the width along the first basis vector is smaller than 2.
4112 This means that there will be at most 2 slices orthogonal to the chosen
4113 direction.
4115 The computation of the above reduced basis requires the solution
4116 of many linear programs, for which we use any of the following
4117 external solvers:
4118 \begin{itemize}
4119 \item \ai[\tt]{GLPK}~\shortcite{GLPK}
4121 This solver is based on double precision floating point arithmetic and
4122 may therefore not be suitable if the coefficients of the constraints
4123 describing the polytope are large.
4125 \item \ai[\tt]{cdd}~\shortcite{cdd}
4127 This solver is based on exact integer arithmetic.
4128 Note that you need version \verb+cddlib 0.94e+ or newer.
4129 Earlier versions (\verb+0.93+--\verb+0.94d+) have
4130 a bug that may sometimes result in a polytope being
4131 reported as (rationally) empty even though it is not.
4133 \item \piplib/~\shortcite{Feautrier:PIP}
4135 This solver is also based on exact integer arithmetic
4136 and uses the \ai{dual simplex} method to solve a linear program.
4137 Two versions are available, \ai[\tt]{pip} will present the
4138 original program to \piplib/, while \ai[\tt]{pip-dual} will present
4139 the dual program to \piplib/, effectively having it apply the primal
4140 simplex method to the original problem.
4141 The latter may seem more appropriate since the computation
4142 of the reduced basis only requires the dual solution of
4143 any linear program. However, in practice, it appears
4144 that \ai[\tt]{pip} is often faster than \ai[\tt]{pip-dual}.
4145 \end{itemize}
4146 The LP solver to use can be selected with the \ai[\tt]{--gbr} option.
4149 \subsection{Computing the integer hull of a polyhedron}
4150 \label{s:integer:hull}
4152 For computing the \ai{integer hull} of a polyhedron,
4153 we first describe how to compute the convex hull of a set
4154 given as an oracle for optimizing a linear objective
4155 function over the set and then
4156 we explain how to optimize a linear objective function over
4157 the integer points of a polyhedron.
4158 Applying the first with the second as \ai{optimization oracle}
4159 yields a method for computing the requested integer hull.
4161 \subsubsection{Computing the convex hull based on an optimization oracle}
4163 The algorithm described below is presented by
4164 \shortciteN[Remark~2.5]{Cook1992} as an extension of the
4165 algorithm by \shortciteN[Section~3]{Edmonds82} for computing
4166 the {\em dimension} of a polytope for which only an optimization oracle
4167 is available. The algorithm is described in a bit more detail
4168 by \shortciteN{Eisenbrand2000PhD} and reportedly stems from
4169 \shortciteN{Hartmann1989PhD}.
4170 Essentially the same algorithm has also been implemented
4171 by \shortciteN{Huggins06}, citing
4172 \ai{beneath/beyond}~\shortcite{Preparata1985} as his inspiration.
4174 The algorithm start out from an initial set of points from
4175 the set $S$. After computing the convex hull of this set
4176 of points, we take one of its bounding constraints and use
4177 the optimization oracle
4178 to compute an optimal point in $S$ (but on the other side
4179 of the bounding hyperplane) along the
4180 outer normal of this bounding constraint.
4181 If a new point is found, it is added to the set of points
4182 and a new convex hull is computed, or the old one is adapted
4183 in a beneath/beyond fashion. Otherwise, the chosen bounding constraint
4184 is also a bounding constraint of $S$ and need not be considered anymore.
4185 The process continues until all bounding constraints in the
4186 description of the current convex hull have been considered.
4188 In principle, the initial set of points in the above algorithm
4189 may be empty, with a ``convex hull'' described by a set of
4190 conflicting constraints and each equality in the description of any
4191 intermediate lower-dimensional convex hull being considered
4192 as a pair of bounding constraints with opposite outer normals.
4193 However, in our implementation, we have chosen to first compute
4194 a maximal set of affinely independent points by first taking any
4195 point from $S$ and then adding points from $S$ not on one of
4196 the equalities satisfied by all points found so far.
4197 This allows us to not have to worry about equalities in the
4198 main algorithm.
4199 In the case of the computation of the integer hull, finding
4200 these affinely independent points can be accomplished using the technique of
4201 \autoref{s:feasibility}.
4203 \begin{figure}
4204 \intercol=0.58cm
4205 \begin{xy}
4206 <\intercol,0pt>:<0pt,\intercol>::
4207 \POS(-1,0)*\xybox{
4208 \def\latticebody{\POS="c"+(0,0.5)\ar@{--}"c"+(0,6.5)}%
4209 \POS0,{\xylattice{1}{6}00}%
4210 \def\latticebody{\POS="c"+(0.5,0)\ar@{--}"c"+(6.5,0)}%
4211 \POS0,{\xylattice00{1}6}%
4212 \POS@i@={(1.5,2.75),(5.75,2.25),(5.5,5.25),(2.75,4.75),(1.5,2.75)},
4213 {0*\xypolyline{}}
4214 \POS@i@={(2,3),(3,3),(3,4),(2,3)},{0*[|(3)]\xypolyline{}}
4215 \POS(2,3)*{\bullet}
4216 \POS(3,3)*{\bullet}
4217 \POS(3,4)*{\bullet}
4218 \POS(3,3.5)\ar(3.5,3.5)
4219 \POS(5,3)*{\circ}
4220 \POS(5,4)*{\circ}
4221 \POS(5,5)*{\circ}
4223 \POS(6,0)*\xybox{
4224 \def\latticebody{\POS="c"+(0,0.5)\ar@{--}"c"+(0,6.5)}%
4225 \POS0,{\xylattice{1}{6}00}%
4226 \def\latticebody{\POS="c"+(0.5,0)\ar@{--}"c"+(6.5,0)}%
4227 \POS0,{\xylattice00{1}6}%
4228 \POS@i@={(1.5,2.75),(5.75,2.25),(5.5,5.25),(2.75,4.75),(1.5,2.75)},
4229 {0*\xypolyline{}}
4230 \POS@i@={(2,3),(5,3),(3,4),(2,3)},{0*[|(3)]\xypolyline{}}
4231 \POS(2,3)*{\bullet}
4232 \POS(5,3)*{\bullet}
4233 \POS(3,4)*{\bullet}
4234 \POS(4,3.5)\ar(4.25,4)
4235 \POS(5,5)*{\circ}
4237 \POS(13,0)*\xybox{
4238 \def\latticebody{\POS="c"+(0,0.5)\ar@{--}"c"+(0,6.5)}%
4239 \POS0,{\xylattice{1}{6}00}%
4240 \def\latticebody{\POS="c"+(0.5,0)\ar@{--}"c"+(6.5,0)}%
4241 \POS0,{\xylattice00{1}6}%
4242 \POS@i@={(1.5,2.75),(5.75,2.25),(5.5,5.25),(2.75,4.75),(1.5,2.75)},
4243 {0*\xypolyline{}}
4244 \POS@i@={(2,3),(5,3),(5,5),(3,4),(2,3)},{0*[|(3)]\xypolyline{}}
4245 \POS(2,3)*{\bullet}
4246 \POS(5,3)*{\bullet}
4247 \POS(5,5)*{\bullet}
4248 \POS(3,4)*{\bullet}
4250 \end{xy}
4251 \caption{The integer hull of a polytope}
4252 \label{f:integer:hull}
4253 \end{figure}
4255 \begin{example}
4256 Assume we want to compute the integer hull of the polytope in the left part
4257 of \autoref{f:integer:hull}.
4258 We first compute a set of three affinely independent points,
4259 shown in the same part of the figure.
4260 Of the three facets of the corresponding convex hull,
4261 optimization along the outer normal (depicted by an arrow in the figure)
4262 of only one facet will yield any additional points. The other two
4263 are therefore facets of the integer hull.
4264 Optimization along the above outer normal may yield any of the
4265 points marked by a $\circ$.
4266 Assuming it is the bottom one, we end up with the updated
4267 convex hull in the middle of the figure. This convex hull
4268 has only one new facet. Adding the point found by optimizing
4269 over this facet's outer normal, we obtain the convex hull
4270 on the right of the figure.
4271 There are two new facets, but neither of them yields any
4272 further points. We have therefore found the integer hull
4273 of the polytope.
4274 \end{example}
4276 \subsubsection{Optimization over the integer points of a polyhedron}
4277 \label{s:optimization}
4279 We assume that we want to find the {\em minimum} of
4280 some linear objective function. When used in the computation
4281 of the integer hull of some polytope, the objective function
4282 will therefore correspond to the inner normal of some facet.
4284 During our search for an optimal integer point with respect
4285 to some objective function, we will keep track of the best
4286 point so far as well as a lower bound $l$
4287 and an upper bound $u$ such that the value at the optimal point
4288 (if it is better than the current best) lies between those
4289 two bounds.
4290 Initially, there is no best point yet and values for $l$ and $u$
4291 may be obtained from optimization over the linear relaxation.
4292 When used in the computation of the integer hull of some polytope,
4293 the upper bound $u$ is one less than the value attained on
4294 the given facet of the current approximation.
4296 As long as $l \le u$, we perform the following steps
4297 \begin{itemize}
4298 \item use the integer feasibility technique of \autoref{s:feasibility}
4299 to test whether there is any integer point with value in
4300 $[l,u']$, where $u'$ is
4301 \begin{itemize}
4302 \item $u$ if the previous test for an integer point did not produce a point
4303 \item $l+\floor{\frac{u-l-1}2}$
4304 if the previous test for an integer point {\em did\/} produce a point
4305 \end{itemize}
4306 \item if a point is found, then remember it as the current best
4307 and replace $u$ by the value at this point minus one,
4308 \item otherwise, replace $l$ by $u'+1$.
4309 \end{itemize}
4310 When used in the computation of the integer hull of some polytope,
4311 it is useful to not only keep track of the best point so far,
4312 but of all points found.
4313 These points will all lie outside of the current approximation
4314 of the integer hull and adding them all instead of just one,
4315 will typically get us to the complete integer hull quicker.
4317 \begin{figure}
4318 \intercol=0.7cm
4319 \begin{xy}
4320 <\intercol,0pt>:<0pt,\intercol>::
4321 \POS(0.5,0)\ar@{-}(16.5,0)
4322 \def\latticebody{\POS="c"+(0,-0.2)\ar@{--}"c"+(0,0.2)\POS"c"*++!D{\the\latticeA}}%
4323 \POS0,{\xylattice{1}{16}00}%
4324 \POS(6,0)*!C{\bullet}
4325 \POS(7,0)*{\bullet}
4326 \POS(8,0)*{\bullet}
4327 \POS(12,0)*{\bullet}
4328 \POS(13,0)*{\bullet}
4329 \POS(14,0)*{\bullet}
4330 \POS(15,0)*{\bullet}
4331 \POS(16,0)*{\bullet}
4332 \POS(1,-1)\ar@{-}(16,-1)
4333 \POS(8,-1)*{\bullet}
4334 \POS(1,-2)\ar@{-}(4,-2)
4335 \POS(5,-3)\ar@{-}(7,-3)
4336 \POS(6,-3)*{\bullet}
4337 \POS(4.9,-4)\ar@{-}(5.1,-4)
4338 \end{xy}
4339 \caption{The integer points of a polytope projected on an objective function}
4340 \label{f:hull:projected}
4341 \end{figure}
4343 \begin{example}
4344 \label{ex:hull:projected}
4345 Assume that the values of some objective function attained
4346 by the integer points of some polytope are as shown in
4347 \autoref{f:hull:projected} and assume we know that the optimal
4348 value lies between 1 and 16.
4349 In the first step we would look for a point attaining a value
4350 in the interval $[1,16]$. Suppose this yields a point attaining
4351 the value $8$ (second line of the figure). We record this point
4352 as the current best and update the search interval to $[1,7]$.
4353 In the second step, we look for a point attaining a value
4354 in the interval $[1,4]$, but find nothing and set the search interval
4355 to $[5,7]$.
4356 In the third step, we consider the interval $[5,7]$ and find
4357 a point attaining the value 6. We update the current best value
4358 and set the search interval to $[5,5]$.
4359 In the fourth step, we consider the interval $[5,5]$, find no
4360 points and update the interval to ``$[6,5]$''.
4361 Since the lower bound is now larger than the upper bound, the
4362 algorithm terminates, returning the best or all point(s) found.
4363 \end{example}
4366 \subsection{Computing the integer hull of a truncated cone}
4367 \label{s:hull:cone}
4369 In \autoref{s:width} we will need to compute the \ai{integer hull}
4370 of a cone with the origin removed ($C \setminus \{ \vec 0 \}$).
4372 \subsubsection{Using the Hilbert basis of the cone}
4374 As proposed by \shortciteN{Koeppe2007personal},
4375 one way of computing this integer hull is to first compute
4376 the \ai{Hilbert basis} of $C$ (see \autoref{s:hilbert})
4377 and to then remove from that Hilbert basis the points that
4378 are not vertices of the integer hull of $C \setminus \{ \vec 0 \}$.
4379 The Hilbert basis of $C$ is the minimal set of points
4380 $\vec b_i \in C \cap \ZZ^d$ such that every integer point
4381 $\vec x \in C \cap \ZZ^d$ can be written as a non-negative
4382 {\em integer} combination of the $\vec b_i$.
4383 The vertices $\vec v_j$ of the integer hull of $C \setminus \{ \vec 0 \}$
4384 are such that every integer point
4385 $\vec x \in (C \cap \ZZ^d) \setminus \{ \vec 0 \}$ can
4386 be written as s non-negative {\em rational} combination of $\vec v_j$.
4387 Clearly, any $\vec v_j$ is also a $\vec b_i$ since $\vec v_j$ can
4388 not be written as the sum of a (rational) convex combination of
4389 other integer points in $(C \cap \ZZ^d) \setminus \{ \vec 0 \}$
4390 and a non-negative combination of the extremal rays $\vec r_k$ of $C$.
4391 A fortiori, it can therefore not be written as an integer combination
4392 of other integer points in $C$.
4393 To obtain the $\vec v_j$ from the $\vec b_i$ we therefore simply
4394 need to remove first $(0,0)$ and then those $\vec b_i$ that are
4395 not an extremal ray and that {\em can} be written as a combination
4397 \vec b_i = \sum_{j \ne i} \vec \alpha_j \vec b_j + \sum_k \beta_k \vec r_k
4398 \qquad\text{with $\alpha_j, \beta_k \ge 0$ and $\sum_{j \ne i} \alpha_j = 1$}
4401 Since the $\vec r_k$ are also among the $\vec b_j$, this can
4402 be simplified to checking whether there exists a rational
4403 solution for $\vec \alpha_j$ to
4405 \vec b_i = \sum_{j \ne i} \vec \alpha_j \vec b_j
4406 \qquad\text{with $\alpha_j \ge 0$ and $\sum_{j \ne i} \alpha_j \ge 1$}
4410 \begin{figure}
4411 \intercol=1.1cm
4412 \begin{xy}
4413 <\intercol,0pt>:<0pt,\intercol>::
4414 \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{*}}
4415 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4416 \POS0,{\xylattice{-0}{5}00}%
4417 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4418 \POS0,{\xylattice00{-4}5}%
4419 \POS0\ar(2,-3)
4420 \POS0\ar(3,4)
4421 \POS(2,-3)*{\bullet}
4422 \POS(3,4)*{\bullet}
4423 \POS(1,1)*{\bullet}
4424 \POS(1,-1)*{\bullet}
4425 \POS(1,0)*{\bullet}
4426 \POS(2,-3)*{\times}
4427 \POS(3,4)*{\times}
4428 \POS(1,1)*{\times}
4429 \POS(1,-1)*{\times}
4430 \end{xy}
4431 \caption{The Hilbert basis and the integer hull of a truncated cone}
4432 \label{f:hilbert:hull}
4433 \end{figure}
4435 \begin{example} \label{ex:hilbert:hull}
4436 Consider the cone
4438 C = \poshull \,\{(2,-3), (3,4)\}
4441 shown in Figure~\ref{f:hilbert:hull}.
4442 The Hilbert basis of this cone is
4443 $$\{(0,0),(2,-3),(3,4),(1,1),(1,-1),(1,0)\}.$$
4444 We have $(1,0) = \frac 1 2 (1,1) + \frac 1 2 (1,-1)$,
4445 while $(1,1)$ and $(1,-1)$ can not be written as
4446 overconvex combinations of the other $\vec b_i \ne \vec 0$.
4447 The vertices of the integer hull of $C \setminus \{ \vec 0 \}$
4448 are therefore
4449 $$\{(2,-3),(3,4),(1,1),(1,-1)\}.$$
4450 \end{example}
4452 \subsubsection{Using generalized basis reduction}
4453 \label{s:hull:cone:gbr}
4455 Another way of computing the integer hull of a truncated cone is to apply
4456 the method of \autoref{s:integer:hull}.
4457 In this case, the initial set of points will consist
4458 of (the smallest integer representatives of) the extremal rays
4459 of the cone, together with the extremal rays themselves.
4460 That is, if $C = \poshull \, \{ \vec r_j \}$ with
4461 $\vec r_j \in \ZZ^d$, then our initial approximation of the
4462 integer hull of $C \setminus \{ \vec 0 \}$ is
4464 \convhull \, \{ \vec r_j \} + \poshull \, \{ \vec r_j \}
4467 Furthermore, we need never consider any
4468 of the bounding constraints that are also bounding constraints
4469 of the original cone.
4470 When optimizing along the normal of any of the other facets, we can
4471 take the lower bound to be $1$. This will ensure that
4472 the origin is excluded, without excluding any other integer points.
4474 \begin{figure}
4475 \intercol=0.5cm
4476 \begin{xy}
4477 <\intercol,0pt>:<0pt,\intercol>::
4478 \POS(0,0)*\xybox{
4479 \POS@i@={(3,-4.5),(2,-3),(3,4),(4.125,5.5),(5.5,5.5),(5.5,-4.5)},{0*[grey]\xypolyline{*}}
4480 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4481 \POS0,{\xylattice{-0}{5}00}%
4482 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4483 \POS0,{\xylattice00{-4}5}%
4484 \POS0\ar(2,-3)
4485 \POS0\ar(3,4)
4486 \POS(2,-3)*{\bullet}
4487 \POS(3,4)*{\bullet}
4488 \POS(1,1)*{\circ}
4490 \POS(8,0)*\xybox{
4491 \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{*}}
4492 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4493 \POS0,{\xylattice{-0}{5}00}%
4494 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4495 \POS0,{\xylattice00{-4}5}%
4496 \POS0\ar(2,-3)
4497 \POS0\ar(3,4)
4498 \POS(2,-3)*{\bullet}
4499 \POS(3,4)*{\bullet}
4500 \POS(1,1)*{\bullet}
4501 \POS(1,-1)*{\circ}
4503 \POS(16,0)*\xybox{
4504 \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{*}}
4505 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4506 \POS0,{\xylattice{-0}{5}00}%
4507 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(5.5,0)}%
4508 \POS0,{\xylattice00{-4}5}%
4509 \POS0\ar(2,-3)
4510 \POS0\ar(3,4)
4511 \POS(2,-3)*{\bullet}
4512 \POS(3,4)*{\bullet}
4513 \POS(1,1)*{\bullet}
4514 \POS(1,-1)*{\bullet}
4516 \end{xy}
4517 \caption{The integer hull of a truncated cone}
4518 \label{f:cone:integer:hull}
4519 \end{figure}
4521 \begin{example}
4522 Consider once more the cone
4524 C = \poshull \,\{(2,-3), (3,4)\}
4526 from Example~\ref{ex:hilbert:hull}.
4527 The initial approximation is
4529 C = \convhull \,\{(2,-3), (3,4)\} + \poshull \,\{(2,-3), (3,4)\}
4532 which is shown on the left of \autoref{f:cone:integer:hull}.
4533 The only bounding constraint that does not correspond to a
4534 bounding constraint of $C$ is $7 x - y \ge 17$.
4535 In the first step, we will therefore look for a point
4536 minimizing $7 x - y$ with values in the interval $[1,16]$.
4537 All values of this objective function in the given interval
4538 attained by points in $C$ are shown in \autoref{f:hull:projected}.
4539 From Example~\ref{ex:hull:projected}, we know that the optimal
4540 value is $6$ and this corresponds to the point $(1,1)$.
4541 Adding this point to our hull, we obtain the approximation
4542 in the middle of \autoref{f:cone:integer:hull}.
4543 This approximation has two new facets.
4544 The bounding constraint $3x - 2 y \ge 1$ will not produce
4545 any new points since we would be looking for one in the
4546 interval ``$[1,0]$''.
4547 The other new bounding constraint is $4x + y \ge 5$.
4548 Minimizing $4 x+ y$ with values in the interval $[1,4]$,
4549 we find the minimal value $3$ corresponding to the point $(1,-1)$.
4550 Adding this point, we obtain the complete integer hull
4551 shown on the right of \autoref{f:cone:integer:hull}.
4552 Note that if in the first step we would have added not only
4553 the point corresponding to the optimal value, but instead
4554 all points found in Example~\ref{ex:hull:projected},
4555 then we would have obtained the complete integer hull directly.
4556 \end{example}
4559 \subsection{Computing the lattice width of a parametric polytope}
4560 \label{s:width}
4562 To compute the \ai{lattice width} of a \ai{parametric polytope},
4563 we essentially use the technique of \shortciteN{Eisenbrand2007parameterised},
4564 which improves upon the technique of \shortciteN{Kannan1992}.
4565 Given a parametric polytope
4567 P(\vec p) = \{\, \vec x \mid A \vec x + \vec b(\vec p) \ge \vec 0 \,\}
4570 the width along a direction $\vec c$ is defined in the same
4571 way as for non-parametric polytopes (see \autoref{s:feasibility}),
4572 \begin{equation}
4573 \label{eq:width}
4574 \width_{\vec c} P(\vec p)
4576 \max \{\, \sp c x \mid \vec x \in P(\vec p) \,\}
4578 \min \{\, \sp c x \mid \vec x \in P(\vec p) \,\}
4580 \end{equation}
4581 The \defindex{lattice width} is the minimum width over all
4582 non-zero integer directions:
4584 \width P(\vec p) =
4585 \min_{\vec c \in \ZZ^d \setminus \{ \vec 0 \} } \width_{\vec c} P(\vec p)
4588 We assume that the parameter domain $Q$ of $P(\vec p)$, i.e., the
4589 set of parameter values for which $P(\vec p) \ne \emptyset$,
4590 is full-dimensional and that for each $\vec p$ from the interior
4591 of $Q$, $P(\vec p)$ is also full-dimensional.
4593 Clearly, for any given direction $\vec c$, the minimum and
4594 maximum in \eqref{eq:width} are attained at (different)
4595 vertices of $P(\vec p)$.
4596 The idea of the algorithm is then to consider all pairs
4597 of parametric vertices of $P(\vec p)$, to compute all candidate
4598 integer directions for a given pair of vertices and then to
4599 compute the minimum width over all candidate integer directions
4600 found.
4602 For any given parametric vertex $\vec v(\vec p)$, the (rational)
4603 directions for which this vertex is minimal can be found as follows.
4604 Let $\vec v(\vec p) + C$ be the \ai{vertex cone} of $\vec v(\vec p)$.
4605 If $\vec v(\vec p)$ is minimal for $\vec c$, then all other points
4606 in the vertex cone must yield a bigger or equal value, i.e.,
4607 $\sp y c \ge 0$ for all $\vec y \in C$.
4608 The set of directions is therefore the \ai{polar cone} $C^*$ of $C$.
4609 Note that, in principle, we should only do this for pairs
4610 of vertices that have a common activity domain, where the
4611 activity domains have been partially opened using the
4612 technique of \autoref{p:inclusion-exclusion} to avoid
4613 multiple vertices that coincide on a lower-dimensional
4614 chamber to all be considered on this intersection.
4615 However, this optimization has currently not been implemented.
4617 Given a pair of vertices $\vec v_1(\vec p)$ and $\vec v_2(\vec p)$,
4618 we may assume that $\vec v_1(\vec p)$ attains the minimum and
4619 $\vec v_2(\vec p)$ attains the maximum.
4620 If $\vec v_1(\vec p) + C_1$ and $\vec v_2(\vec p) + C_2$ are the
4621 corresponding vertex cones, then the set of (rational) directions for this
4622 pair of vertices is
4624 C_{1,2} = \left( C_1^* \cap -C_2^* \right) \setminus \{ \vec 0 \}
4627 The set of candidate integer directions are therefore
4628 the vertices of the integer hull of $C_{1,2}$, which
4629 can be computed as explained in \autoref{s:hull:cone}.
4630 To see this, note that by construction
4631 $\sps {\vec c}{\vec v_1(\vec p)} \le \sps {\vec c}{\vec v_2(\vec p)}$
4632 and so
4634 w_{\vec c}(\vec p) = \width_{\vec c} P(\vec p)
4635 = \sps {\vec c}{\vec v_2(\vec p)-\vec v_1(\vec p)} \ge 0
4638 Any integer direction in $C_{1,2}$ will therefore yield
4639 a width that is at least as large as that of one
4640 of the vertices of the integer hull.
4641 Note that when using generalized basis reduction
4642 to compute the integer hull of these cones as in \autoref{s:hull:cone:gbr},
4643 it can be helpful to use as vertices for the initial approximation
4644 not only the extremal rays of the cone, but also those vertices
4645 of previously computed integer hulls that are elements of the current cone.
4647 After computing a list of all possible candidate width directions
4648 $\vec c_i$ and the corresponding widths $w_{\vec c_i}(\vec p)$,
4649 we keep only a single direction of all those that yield
4650 the same width (as an affine function of the parameters).
4651 Then we construct the chambers where each of the widths is minimal,
4652 i.e.,
4654 C_i = \{\, \vec p \in Q \mid \forall j :
4655 w_{\vec c_i}(\vec p) \le w_{\vec c_j}(\vec p) \,\}
4658 Note that many of the $C_i$ may be empty or of lower dimension
4659 than Q and that the other $C_i$ will intersect in common facets.
4660 To obtain a partition of partially-open full-dimensional chambers, we proceed
4661 as in \autoref{s:triangulation}.
4663 \begin{figure}
4664 \intercol=1.1cm
4665 \begin{xy}
4666 <\intercol,0pt>:<0pt,\intercol>::
4667 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
4668 \POS0,{\xylattice{-0}{10}00}%
4669 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(10.5,0)}%
4670 \POS0,{\xylattice00{-0}7}%
4671 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*[|(2)]\xypolyline{}}
4672 \POS(0,0)*{\bullet}
4673 \POS(5,3)*{\bullet}
4674 \POS(5,4)*{\bullet}
4675 \POS(9,6)*{\bullet}
4676 \POS(3,2)*{\bullet}
4677 \POS(4,3)*{\bullet}
4678 \POS(6,4)*{\bullet}
4679 \POS(7,5)*{\bullet}
4680 \POS(9,6);(8.7,6.4)**{}?(0)/1.1cm/="a"\POS(9,6)\ar"a"
4681 \POS(9,6);(9.1,5.8)**{}?(0)/1.1cm/="a"\POS(9,6)\ar"a"
4682 \POS(5,4);(5.4,3.5)**{}?(0)/1.1cm/="a"\POS(5,4)\ar"a"
4683 \POS(5,4);(5.1,3.8)**{}?(0)/1.1cm/="a"\POS(5,4)\ar"a"
4684 \POS(0,0);(0.4,-0.5)**{}?(0)/1.1cm/="a"\POS(0,0)\ar"a"
4685 \POS(0,0);(-0.3,0.5)**{}?(0)/1.1cm/="a"\POS(0,0)\ar"a"
4686 \POS(5,3);(4.7,3.5)**{}?(0)/1.1cm/="a"\POS(5,3)\ar"a"
4687 \POS(5,3);(4.7,3.4)**{}?(0)/1.1cm/="a"\POS(5,3)\ar"a"
4688 \POS(9,6)*+!DL{\vec v_1}
4689 \POS(0,0)*+!UR{\vec v_3}
4690 \POS(5,3)*+!UL{\vec v_4}
4691 \POS(5,4)*+!DR{\vec v_2}
4692 \end{xy}
4693 \caption{A polytope and its candidate width directions}
4694 \label{f:width}
4695 \end{figure}
4697 \begin{example} \label{ex:width}
4698 Consider the (non-parametric) polytope
4700 P = \left\{\,
4701 \vec x \mid
4702 \begin{aligned}
4703 -3 x_1 +5 x_2 &\ge 0 \\
4704 4 x_1 -5 x_2 &\ge 0 \\
4705 x_1 -2 x_2 + 3 &\ge 0 \\
4706 -3 x_1 +4 x_2 + 3 &\ge 0
4707 \end{aligned}
4708 \,\right\}
4710 shown in \autoref{f:width}. The polytope has four vertices
4712 \begin{aligned}
4713 \vec v_1 & = (9,6) \\
4714 \vec v_2 & = (5,4) \\
4715 \vec v_3 & = (0,0) \\
4716 \vec v_4 & = (5,3)
4718 \end{aligned}
4720 The corresponding cones of directions (for
4721 the given vertex to attain the minimum), also shown
4722 in \autoref{f:width} are
4724 \begin{aligned}
4725 C^*_1 & = \poshull \,\{ (-3,4), (1,-2) \} \\
4726 C^*_2 & = \poshull \,\{ (4,-5), (1,-2) \} \\
4727 C^*_3 & = \poshull \,\{ (4,-5), (-3,5) \} \\
4728 C^*_4 & = \poshull \,\{ (-3,5), (-3,4) \}
4730 \end{aligned}
4733 \begin{figure}
4734 \intercol=0.8cm
4735 \begin{xy}
4736 <\intercol,0pt>:<0pt,\intercol>::
4737 \def\latticebody{\POS="c"+(0,-6.5)\ar@{--}"c"+(0,2.5)}%
4738 \POS0,{\xylattice{-1}{5}00}%
4739 \def\latticebody{\POS="c"+(-1.5,0)\ar@{--}"c"+(5.5,0)}%
4740 \POS0,{\xylattice00{-6}2}%
4741 \POS0\ar@{->}(3,-4)\POS?!{(0,-6.5);(1,-6.5)}="a"
4742 \POS0\ar@{->}(-1,2)
4743 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4744 \POS0\ar@{->}(1,-2)
4745 \POS@i@={"a",(3,-4),(4,-5),"b"},{0*[grey]\xypolyline{*}}
4746 \POS0,{\ellipse(1.1)(*0;(4,3)*),^,(*0;(-2,-1)*){-}}
4747 \POS0,{\ellipse(1)(*0;(2,1)*),^,(*0;(5,4)*){-}}
4748 \POS0\ar@{->}(3,-4)\POS?!{(0,-6.5);(1,-6.5)}="a"
4749 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4750 \POS(4,-5)*{\bullet}
4751 \POS(3,-4)*{\bullet}
4752 \end{xy}
4753 \caption{The cone of directions $C_{2,1}$}
4754 \label{f:C:2:1}
4755 \end{figure}
4757 \begin{figure}
4758 \intercol=0.8cm
4759 \begin{xy}
4760 <\intercol,0pt>:<0pt,\intercol>::
4761 \def\latticebody{\POS="c"+(0,-6.5)\ar@{--}"c"+(0,5.5)}%
4762 \POS0,{\xylattice{-3}{5}00}%
4763 \def\latticebody{\POS="c"+(-3.5,0)\ar@{--}"c"+(5.5,0)}%
4764 \POS0,{\xylattice00{-6}5}%
4765 \POS0\ar@{->}(3,-4)
4766 \POS0\ar@{->}(-1,2)\POS?!{(0,5.5);(1,5.5)}="a"
4767 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4768 \POS0\ar@{->}(-3,5)
4769 \POS@i@={"b",(4,-5),(1,-1),(-1,2),"a",(5.5,5.5),(5.5,-6.5)},{0*[grey]\xypolyline{*}}
4770 \POS0\ar@{->}(-1,2)\POS?!{(0,5.5);(1,5.5)}="a"
4771 \POS0\ar@{->}(4,-5)\POS?!{(0,-6.5);(1,-6.5)}="b"
4772 \POS0,{\ellipse(1.1)(*0;(4,3)*),^,(*0;(-2,-1)*){-}}
4773 \POS0,{\ellipse(1)(*0;(5,4)*),^,(*0;(-5,-3)*){-}}
4774 \POS(1,-1)*{\bullet}
4775 \POS(4,-5)*{\bullet}
4776 \POS(-1,2)*{\bullet}
4777 \end{xy}
4778 \caption{The cone of directions $C_{3,1}$}
4779 \label{f:C:3:1}
4780 \end{figure}
4782 \begin{figure}
4783 \intercol=0.8cm
4784 \begin{xy}
4785 <\intercol,0pt>:<0pt,\intercol>::
4786 \def\latticebody{\POS="c"+(0,-4.5)\ar@{--}"c"+(0,5.5)}%
4787 \POS0,{\xylattice{-3}{3}00}%
4788 \def\latticebody{\POS="c"+(-3.5,0)\ar@{--}"c"+(3.5,0)}%
4789 \POS0,{\xylattice00{-4}5}%
4790 \POS0\ar@{->}(3,-4)
4791 \POS0\ar@{->}(-1,2)
4792 \POS0,{\ellipse(1.1)(*0;(4,3)*),^,(*0;(-2,-1)*){-}}
4793 \POS0\ar@{->}(-3,5)
4794 \POS0\ar@{->}(-3,4)
4795 \POS0,{\ellipse(1)(*0;(-5,-3)*),^,(*0;(-4,-3)*){-}}
4796 \end{xy}
4797 \caption{The cone of directions $C_{4,1}$}
4798 \label{f:C:4:1}
4799 \end{figure}
4801 Let us now consider the directions in which
4802 $\vec v_2$ is minimal while $\vec v_1$ is maximal.
4803 We find
4805 C_{2,1} = \poshull \,\{ (4,-5), (3,-4) \} \setminus \{ \vec 0 \}
4808 as shown in \autoref{f:C:2:1}.
4809 The vertices of the integer hull of $C_{2,1}$ are $(4,-5)$
4810 and $(3,-4)$.
4811 The corresponding widths are
4813 \begin{aligned}
4814 \vec c_1 &= (4,-5) & w_{\vec c_1} &= 6 \\
4815 \vec c_2 &= (3,-4) & w_{\vec c_2} &= 4
4817 \end{aligned}
4819 We similarly find
4821 C_{3,1} = \poshull \,\{ (4,-5), (-1,2) \} \setminus \{ \vec 0 \}
4824 with integer hull
4825 $\poshull \,\{ (4,-5), (-1,2), (1,-1) \}$, shown
4826 in \autoref{f:C:3:1}, yielding
4828 \begin{aligned}
4829 \vec c_3 &= (4,-5) & w_{\vec c_3} &= 6 \\
4830 \vec c_4 &= (-1,2) & w_{\vec c_4} &= 3 \\
4831 \vec c_5 &= (1,-1) & w_{\vec c_5} &= 3
4833 \end{aligned}
4835 On the other hand,
4837 C_{4,1} = \emptyset
4840 as shown in \autoref{f:C:4:1} and so this combination
4841 does not yield any width direction candidates.
4842 The other pairs of vertices further yield
4844 \begin{aligned}
4845 \vec c_6 &= (-1,2) & w_{\vec c_6} &= 3 \\
4846 \vec c_7 &= (-3,5) & w_{\vec c_7} &= 5 \\
4847 \vec c_8 &= (-3,4) & w_{\vec c_8} &= 4 \\
4848 \vec c_9 &= (-3,5) & w_{\vec c_9} &= 5 \\
4849 \vec c_{10} &= (-2,3) & w_{\vec c_{10}} &= 3
4851 \end{aligned}
4853 Since the polytope under consideration is not parametric,
4854 there is only one (non-empty, $0$-dimensional) chamber and
4855 it corresponds to one of the directions, say $\vec c_4 = (-1,2)$,
4856 with width $3$ (the other directions with the same width
4857 having been removed).
4859 \begin{figure}
4860 \intercol=1.1cm
4861 \begin{xy}
4862 <\intercol,0pt>:<0pt,\intercol>::
4863 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
4864 \POS0,{\xylattice{-0}{10}00}%
4865 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(10.5,0)}%
4866 \POS0,{\xylattice00{-0}7}%
4867 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*[|(2)]\xypolyline{}}
4868 \POS(-0.5,-0.5)\ar@{.}(7.5,7.5)
4869 \POS(0.5,-0.5)\ar@{.}(8.5,7.5)
4870 \POS(1.5,-0.5)\ar@{.}(9.5,7.5)
4871 \POS(2.5,-0.5)\ar@{.}(10.5,7.5)
4872 \POS(-0.5,-0.25)\ar@{-}(10.5,5.25)
4873 \POS(-0.5,0.25)\ar@{-}(10.5,5.75)
4874 \POS(-0.5,0.75)\ar@{-}(10.5,6.25)
4875 \POS(-0.5,1.25)\ar@{-}(10.5,6.75)
4876 \POS(-0.25,-0.5)\ar@{--}(10.5,6.666)
4877 \POS(-0.5,-0.333)\ar@{--}(10.5,7)
4878 \POS(-0.5,0)\ar@{--}(10.5,7.333)
4879 \POS(-0.5,0.333)\ar@{--}(10.25,7.5)
4880 \POS(0,0)*{\bullet}
4881 \POS(5,3)*{\bullet}
4882 \POS(5,4)*{\bullet}
4883 \POS(9,6)*{\bullet}
4884 \POS(3,2)*{\bullet}
4885 \POS(4,3)*{\bullet}
4886 \POS(6,4)*{\bullet}
4887 \POS(7,5)*{\bullet}
4888 \end{xy}
4889 \caption{A polytope and its lattice width directions}
4890 \label{f:width:2}
4891 \end{figure}
4893 Each of the three directions that yield the minimal
4894 width of 3 is shown in \autoref{f:width:2}.
4895 \end{example}
4897 \begin{example} \label{ex:width:2}
4898 Consider the polytope
4900 P(p) = \left\{\,
4901 \vec x \mid
4902 \begin{aligned}
4903 -2 x_1 + p + 5 &\ge 0 \\
4904 2 x_1 + p + 5 &\ge 0 \\
4905 -2 x_2 - p + 5 &\ge 0 \\
4906 2 x_2 - p + 5 &\ge 0
4907 \end{aligned}
4908 \,\right\}
4910 from \shortciteN[Example~2.1.7]{Woods2004PhD}.
4911 The parametric vertices are
4913 \begin{aligned}
4914 \vec v_1(p) & = \left(\frac{p+5}2, \frac{-p+5}2\right) \\
4915 \vec v_2(p) & = \left(\frac{p+5}2, \frac{p-5}2\right) \\
4916 \vec v_3(p) & = \left(\frac{-p-5}2, \frac{-p+5}2\right) \\
4917 \vec v_4(p) & = \left(\frac{-p-5}2, \frac{p-5}2\right)
4919 \end{aligned}
4921 We find two essentially different candidate width directions
4923 \begin{aligned}
4924 \vec c_1 &= (0,1) & w_{\vec c_1}(p) &= 5-p \\
4925 \vec c_2 &= (1,0) & w_{\vec c_2}(p) &= 5+p
4927 \end{aligned}
4929 The first direction can be found by combining, say,
4930 $\vec v_1(p)$ and $\vec v_2(p)$, while the second direction can be
4931 found by combining, say, $\vec v_1(p)$ and $\vec v_3(p)$.
4932 The parameter domain for the parametric polytope $P(p)$ is
4934 Q = \{\, p \mid -5 \le p \le 5 \,\}
4937 The two (closed) chambers are therefore
4939 \begin{aligned}
4940 C_1 &= \{\, p \in Q \mid 5 - p \le 5+p \,\} \\
4941 C_2 &= \{\, p \in Q \mid 5 + p \le 5-p \,\}
4943 \end{aligned}
4945 To obtain a partition, \autoref{s:interior} gives
4946 the internal point $(0,0)$, which happens to meet
4947 the facets $p \ge 0$ and $-p \ge 0$. We therefore
4948 keep the facet with positive (inner) normal closed
4949 and open up the other facet. The result is
4951 \begin{aligned}
4952 \hat C_1 &= \{\, p \mid 0 \le p \le 5 \,\} \\
4953 \hat C_2 &= \{\, p \mid -5 \le p < 0 \,\}
4955 \end{aligned}
4957 Since we are usually only interested in integer parameter
4958 values, the latter chamber would become
4959 $\hat C_2 = \{\, p \mid -5 \le p \le -1 \,\}$.
4960 \end{example}
4962 Our description differs slightly from that of
4963 of \shortciteN{Eisenbrand2007parameterised}.
4964 First, they consider all pairs of basic solutions instead
4965 of all pairs of vertices, which means that they may
4966 consider basic solutions that are never feasible and that,
4967 in case of a non-simple polytope,
4968 they may consider the same parametric vertex more than once.
4969 The set of integer
4970 directions for a pair of vertices is the intersection of
4971 the sets of integer directions they obtain for each of
4972 the corresponding basic solutions.
4973 Second, they use a different method of creating a partition
4974 of partially-open chambers, which may lead to some lower-dimensional
4975 chambers surviving and hence to a larger total number of chambers.
4978 \subsection{Testing whether a set has an infinite number of points}
4979 \label{s:infinite}
4981 In some situation we are given the generating function of
4982 some integer set and we would like to know if the set is
4983 infinite or not. Typically, we want to know if the set
4984 is empty or not, but we cannot simply count the number of elements
4985 in the standard way since we may not have any guarantee that
4986 the set has only a finite number of elements.
4987 We will consider the slightly more general case where we are
4988 given a rational generating function $f(\vec x)$ of the form~\eqref{eq:rgf}
4989 such that
4990 \begin{equation}
4991 \label{eq:rgf:psp}
4992 f(\vec x) = \sum_{\vec s \in Q \cap \ZZ^d} c(\vec s)\, \vec x^{\vec s}
4993 \end{equation}
4994 converges on some nonempty open subset of $\CC^d$, $Q$ is a pointed
4995 polyhedron and $c(\vec s) \ge 0$,
4996 and we want to compute
4997 \begin{equation}
4998 \label{eq:psp:sum}
4999 S = \sum_{\vec s \in Q \cap \ZZ^d} c(\vec s)
5001 \end{equation}
5002 where the sum may diverge, i.e., ``$S = \infty$''.
5003 The following proposition shows that we can determine $S$
5004 in polynomial time.
5005 For a sketch of an alternative technique, see
5006 \shortciteN[Proof of Lemma~16]{Woods2005period}.
5008 \begin{proposition}
5009 Fix $d$ and $k$.
5010 Given a \rgf/ of the form~\eqref{eq:rgf} with $k_i \le k$
5011 and a pointed polyhedron $Q \subset \QQ^d$, then there is a
5012 polynomial time algorithm that determines for the corresponding
5013 function $c(\vec s)$~\eqref{eq:rgf:psp} whether the sum~\eqref{eq:psp:sum}
5014 diverges and computes the value of $S$~\eqref{eq:psp:sum} if it does not.
5015 \end{proposition}
5016 \begin{proof}
5017 Since $Q$ is pointed, the series~\eqref{eq:rgf:psp} converges on a neighborhood
5018 of $e^{\vec \ell} = (e^{\ell_1}, \ldots, e^{\ell_d})$ for any $\vec \ell$
5019 such that $\sps {\vec r_k} {\vec \ell} < 0$ for
5020 any (extremal) ray $\vec r_k$ of $Q$ and
5021 such that $\sps {\vec b_{i j}} {\vec \ell} \ne 0$ for any
5022 $\vec b_{i j}$ in~\eqref{eq:rgf}.
5023 Let $\vec \alpha = - \vec \ell$ and perform the substitution
5024 $\vec x = t^{\vec \alpha}$. The function $g(t) = f(t^{\vec \alpha})$
5025 is then also a (short) \rgf/ and
5027 g(t) = \sum_{k \in \sps {\vec\alpha} Q \cap \ZZ}
5028 \left(
5029 \sum_{\shortstack{$\scriptstyle \vec s \in Q \cap \ZZ^d$\\
5030 $\scriptstyle \sp \alpha s = k$}} c(\vec s)
5031 \right) t^k
5032 =: \sum_{k \in \sps {\vec\alpha} Q \cap \ZZ} d(k) \, t^k
5035 with $\sps {\vec\alpha} Q = \{ \sp \alpha x \mid \vec x \in Q \}$,
5036 converges in a neighborhood of $e^{-1}$, while
5038 S = \sum_{k \in \sps {\vec\alpha} Q \cap \ZZ} d(k)
5041 Since $c(\vec s) \ge 0$, we have $d(k) \ge 0$
5042 and the above sum diverges iff any of the coefficients of the
5043 negative powers of $t$ in the Laurent expansion of $g(t)$ is non-zero.
5044 If the sum converges, then the sum is simply the coefficient
5045 of the constant term in this expansion.
5047 It only remains to show now that we can compute a suitable $\vec \alpha$
5048 in polynomial time, i.e., an $\vec \alpha$ such that
5049 $\sps {\vec r_k} {\vec \alpha} > 0$ for any (extremal) ray $\vec r_k$ of $Q$ and
5050 $\sps {\vec b_{i j}} {\vec \alpha} \ne 0$ for any
5051 $\vec b_{i j}$ in~\eqref{eq:rgf}.
5052 By adding the $\vec r_k$ to the list of $\vec b_{i j}$ if needed, we can relax
5053 the first set of constraints to $\sps {\vec r_k} {\vec \alpha} \ge 0$.
5054 Let $Q$ be described by the constraints $A \vec x + \vec c \ge \vec 0$
5055 and let $B$ be $d \times d$ non-singular submatrix of $A$, obtained
5056 by removing some of the rows of $A$. Such a $B$ exists since
5057 $Q$ does not contain any straight line.
5058 Clearly, $B \vec r \ge \vec 0$ for any ray $\vec r$ of $Q$.
5059 Let $\vec b'_{i j} = B \vec b_{i j}$, then since $\vec b_{i j} \ne \vec 0$
5060 and B is non-singular, we have $\vec b'_{i j} \ne \vec 0$.
5061 We may therefore find in polynomial time a point $\vec \alpha' \ge \vec 0$
5062 on the ``\ai{moment curve}'' such that
5063 $\sps {\vec b'_{i j}} {\vec \alpha'} \ne 0$
5064 \shortcite[Algorithm~5.2]{Barvinok1999}.
5065 Let $\vec \alpha = B^\T \vec \alpha'$.
5066 Then
5068 \sps {\vec b_{i j}} {\vec \alpha}
5070 \sps {\vec b_{i j}} {B^\T \vec \alpha'}
5072 \sps {B \vec b_{i j}} {\vec \alpha'}
5074 \sps {\vec b'_{i j}} {\vec \alpha'}
5075 \ne 0
5079 \sps {\vec r_k} {\vec \alpha}
5081 \sps {\vec r_k} {B^\T \vec \alpha'}
5083 \sps {B \vec r_k} {\vec \alpha'}
5084 \ge 0
5087 as required.
5088 Note that in practice, we would, as usual, first try a
5089 fixed number of random vectors $\vec \alpha' \ge \vec 0$
5090 before resorting to looking for a point on the moment curve.
5091 \end{proof}
5094 \subsection{Enumerating integer projections of parametric polytopes}
5095 \label{s:projection}
5097 In this section we are interested in computing
5098 \begin{equation}
5099 \label{eq:count:projection}
5100 c(\vec s)=\#\left\{\vec t\in\ZZ^{d} \mid \exists \vec u\in\ZZ^{m}:
5101 (\vec s,\vec t,\vec u)\in P\right\}
5103 \end{equation}
5104 with $P \subset \QQ^{n}\times\QQ^{d}\times\QQ^{m}$ a rational
5105 pointed polyhedron such that
5107 P_{\vec s}=\left\{(\vec t,\vec u)\in\QQ^d\times\QQ^m
5108 \mid (\vec s,\vec t,\vec u)\in P\right\}
5110 is a polytope for any $\vec s$.
5111 This is equivalent to computing the number of points
5112 in the \ai{integer projection} of a parametric polytope
5114 c(\vec s)=\#\big(\pi(P_{\vec s}\cap\ZZ^{d+m})\big)
5117 with $\pi:\QQ^d\times\QQ^m\rightarrow\QQ^d$ defined by
5118 $\pi(\vec t, \vec u)=\vec t$.
5119 Exponential methods for computing $c(\vec s)$ are
5120 described by \shortciteN{Verdoolaege2005experiences}
5121 and \shortciteN{Seghir2006memory}.
5122 Here, we provide some implementation details for the polynomial
5123 method of \shortciteN[Theorem~1.7]{Woods2003short}, for
5124 computing the generating function, $\sum_{\vec s}c(\vec s) \, \vec x^{\vec s}$,
5125 which can then be converted into an explicit function $c(\vec s)$
5126 \shortcite[Corollary~1.11]{Verdoolaege2008counting}.
5127 Note that in contrast to \shortciteN[Theorem~1.7]{Woods2003short},
5128 we may allow $P$ to be an unbounded (but still pointed) polyhedron here
5129 (as long as $P_{\vec s}$ is bounded), since
5130 we replace their application of
5131 \shortciteN[Lemma~3.1]{Kannan1992}
5132 by \shortciteN[Theorem~5]{Eisenbrand2007parameterised}.
5134 If there is only one existentially quantified variable ($m = 1$),
5135 then computing~\eqref{eq:count:projection} is easy.
5136 You simply shift $P$ by $1$ in the $u$ direction and subtract
5137 this shifted copy from the original,
5139 D = P \setminus (\vec e_{n+d+1} + P)
5142 (See, e.g., \shortciteN[Figure~1, page~973]{Woods2003short}
5143 or \shortciteN[Figure~4.33, page~186]{Verdoolaege2005PhD}.)
5144 In the difference $D$ there will be {\em exactly} one value of $u$
5145 for each value of the remaining variables for which there was
5146 {\em at least} one value of $u$ in $P$,
5148 \forall (\vec s, \vec t):
5149 \quad
5150 \left(
5151 \exists u: (\vec s, \vec t, u) \in P
5152 \right)
5153 \iff
5154 \left(
5155 \exists! u: (\vec s, \vec t, u) \in D
5156 \right)
5159 The function $c(\vec s)$ can then be computed by counting
5160 the number of elements in $D(\vec s)$.
5161 These operations can be performed either in the space
5162 of (unions of) parametric polytopes or on generating functions.
5163 In the first case, $D(\vec s)$ can be written as a disjoint union
5164 of parametric polytopes that can be enumerated separately.
5165 In the second case, we first compute the generating function
5166 $f(\vec x, \vec y)$ of the set
5170 (\vec s, \vec t) \mid \exists u \in \ZZ : (\vec s, \vec t, u) \in P
5173 and then obtain the generating function $C(\vec x)$ of $c(\vec s)$
5174 as $C(\vec x) = f(\vec x, \vec 1)$.
5175 In the remainder of this section, we will concentrate on the
5176 computation of the generating function of $S$.
5177 To compute this generating function in the current case where
5178 there is only one existentially quantified variable, we first
5179 compute the generating function $g(\vec x, \vec y, z)$ of
5180 $P(\vec s, \vec t, u)$, perform operations on the generating function
5181 equivalent to the set operations (see, e.g.,
5182 \shortciteN[Section~4.5.3]{Verdoolaege2005PhD}), resulting
5183 in a generating function $g'(\vec x, \vec y, z)$, and then sum
5184 over all values (at most one for each value of $\vec s$
5185 and $\vec t$) of $u$, i.e., $f(\vec x, \vec y) = g'(\vec c, \vec y, 1)$.
5187 \begin{figure}
5188 \intercol=1.05cm
5189 \begin{xy}
5190 <\intercol,0pt>:<0pt,\intercol>::
5191 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
5192 \POS0,{\xylattice{-0}{10}00}%
5193 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(10.5,0)}%
5194 \POS0,{\xylattice00{-0}7}%
5195 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*\xypolyline{}}
5196 \POS(0,0)*[*0.333]\xybox{
5197 \POS@i@={(0,0),(5,3),(9,6),(5,4),(0,0)},{0*\xypolyline{--}}
5199 \POS(0,0)*{\bullet}
5200 \POS(5,3)*{\bullet}
5201 \POS(5,4)*{\bullet}
5202 \POS(9,6)*{\bullet}
5203 \POS(3,2)*{\bullet}
5204 \POS(4,3)*{\bullet}
5205 \POS(6,4)*{\bullet}
5206 \POS(7,5)*{\bullet}
5207 \POS(-1,-0.5)\ar@{-}(-1,7.5)
5208 \POS(-1,0)*{\bullet}
5209 \POS(-1,3)*{\bullet}
5210 \POS(-1,4)*{\bullet}
5211 \POS(-1,6)*{\bullet}
5212 \POS(-1,2)*{\bullet}
5213 \POS(-1,3)*{\bullet}
5214 \POS(-1,4)*{\bullet}
5215 \POS(-1,5)*{\bullet}
5216 \POS(-0.5,-1)\ar@{-}(10.5,-1)
5217 \POS(0,-1)*+++!UR{S}
5218 \POS(0,-1)*{\bullet}
5219 \POS(5,-1)*{\bullet}
5220 \POS(5,-1)*{\bullet}
5221 \POS(9,-1)*{\bullet}
5222 \POS(3,-1)*{\bullet}
5223 \POS(4,-1)*{\bullet}
5224 \POS(6,-1)*{\bullet}
5225 \POS(7,-1)*{\bullet}
5226 \end{xy}
5227 \caption{A polytope and its integer projections}
5228 \label{f:projection}
5229 \end{figure}
5231 If there is more than one existentially quantified variable ($m > 1$),
5232 then we can in principle apply the above shifting and subtracting
5233 technique recursively to obtain a generating function
5234 $f(\vec x, \vec y)$ for the set
5235 \begin{equation}
5236 \label{eq:projection:T}
5239 (\vec s, \vec t) \mid \exists \vec u \in \ZZ^m : (\vec s, \vec t, \vec u) \in P
5241 \end{equation}
5242 and then compute $C(\vec x) = f(\vec x, \vec 1)$.
5243 There are however some complications.
5244 Most notably, after applying the technique in one direction
5245 and projecting out the corresponding variable, the resulting set, i.e.,
5249 (\vec s, \vec t, u_1, \ldots, u_{m-1}) \mid
5250 \exists u_m \in \ZZ : (\vec s, \vec t, \vec u) \in P
5254 in general does not correspond to the integer points in some polytope.
5255 For example, assume that the polytope in \autoref{f:projection}
5256 contains the values of $\vec u$ associated to a particular
5257 value of $(\vec s, \vec t)$. Since there are integer points
5258 in this polytope, we should count this value of $\vec t$, but only once.
5259 If we apply the above technique in the vertical direction ($u_2$), then
5260 we can compute (a generating function for) the set $S$ shown
5261 on the bottom of the figure.
5262 Note, however, that there are ``gaps'' in this set, i.e.,
5263 if we compute $S \setminus (\vec e_{n+d+1} + S)$ then we will not
5264 end up with a single point (for this value of $(\vec s, \vec t)$).
5265 Since the biggest gap is three wide, we need
5266 to compute
5269 \setminus (\vec e_{n+d+1} + S)
5270 \setminus (2 \vec e_{n+d+1} + S)
5271 \setminus (3 \vec e_{n+d+1} + S)
5273 to obtain a single point.
5274 If we do the subtraction in the horizontal direction first,
5275 then we end up with a set (shown on the left) with gaps
5276 at most two wide, so afterwards we only need to subtract twice in the
5277 vertical direction.
5279 \begin{figure}
5280 \intercol=0.8cm
5281 \begin{xy}
5282 <\intercol,0pt>:<0pt,\intercol>::
5283 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,7.5)}%
5284 \POS0,{\xylattice{-0}{4}00}%
5285 \def\latticebody{\POS="c"+(-0.5,0)\ar@{--}"c"+(4.5,0)}%
5286 \POS0,{\xylattice00{-0}7}%
5287 \POS@i@={(0,0),(1,3),(3,6),(3,4),(0,0)},{0*\xypolyline{}}
5288 \POS(0,0)*{\bullet}
5289 \POS(1,3)*{\bullet}
5290 \POS(3,4)*{\bullet}
5291 \POS(3,6)*{\bullet}
5292 \POS(1,2)*{\bullet}
5293 \POS(2,3)*{\bullet}
5294 \POS(2,4)*{\bullet}
5295 \POS(3,5)*{\bullet}
5296 \POS(-0.5,-1)\ar@{-}(4.5,-1)
5297 \POS(0,-1)*+++!UR{S}
5298 \POS(0,-1)*{\bullet}
5299 \POS(1,-1)*{\bullet}
5300 \POS(2,-1)*{\bullet}
5301 \POS(3,-1)*{\bullet}
5302 \end{xy}
5303 \caption{A transformed polytope and its integer projection}
5304 \label{f:projection:2}
5305 \end{figure}
5307 In general, there is no bound on the widths of the gaps we may
5308 encounter in any given direction. However, there are directions
5309 in which the gaps are known to be ``small''.
5310 In particular, if the dimension $m$ is fixed, then the lattice width
5311 (see \autoref{s:width}) of lattice point free polytopes
5312 is bounded by a constant $\omega(m)$%
5313 ~\shortcite{Lagarias90,Barvinok02,Banaszczyk1999flatness}.
5314 This means that in the direction of the lattice width of a polytope,
5315 the gaps will be not be larger than $\omega(m)$
5316 \shortcite[Theorem~4.3]{Woods2003short}.
5317 Otherwise, we would be able to put a (uniformly) scaled down version
5318 of the polytope in the gap and it would contain no lattice points,
5319 which would contradict the fact that its lattice width is bounded
5320 by $\omega(m)$.
5321 \autoref{f:projection} contains such a scaled down copy
5322 of the original polytope. However, neither the horizontal
5323 nor the vertical direction is a lattice width direction
5324 of this polytope. The actual lattice width of this
5325 polytope was computed in Example~\ref{ex:width} as $3$
5326 with corresponding direction $\vec c = (-1,2)$.
5327 \autoref{f:projection:2} shows the result of applying
5328 the unimodular transformation
5330 \begin{bmatrix}
5331 -1 & 2 \\
5332 0 & 1
5333 \end{bmatrix}
5335 to the polytope. Note that the horizontal direction
5336 now has gaps of width at most 1. After shifting, subtracting
5337 and projecting in the vertical direction, we therefore
5338 end up with a set $S$ with gaps of width 1 and we then
5339 only have to shift and subtract once in the remaining
5340 (horizontal) direction.
5342 In fact, for two-dimensional polytopes the gaps in the lattice
5343 width direction will always be one, as shown by the following lemma.
5344 \begin{lemma}
5345 \label{l:gap}
5346 For any rational polygon, the gaps in a lattice
5347 width direction are of width at most 1.
5348 \end{lemma}
5349 \begin{proof}
5350 We may assume that $x$ is the given lattice width direction of
5351 a given polygon $P$.
5352 If there is a gap of width 2, then there is an integer value $x_1$ of $x$
5353 such that
5354 $P \cap \{\, (x_1, y) \,\} \ne \emptyset$,
5355 $P \cap \{\, (x_1+2, y) \,\} \ne \emptyset$,
5356 while
5357 $P \cap \{\, (x_1+1, y) \,\} \cap \ZZ^2 = \emptyset$.
5358 Using \shortciteN[Lemma~4.2]{Woods2003short}, we can put
5359 a scaled down copy $P'$ of $P$ between $x=x_1$ and $x=x_1+2$
5360 (and inside of $P$).
5361 $P'$ meets the line $x=x_1+1$ between two consecutive integer
5362 points, $y_1$ and $y_1+1$. Let $P''$ be the polygon bounded by $x=x_1$ and
5363 $x=x_1+2$ and two lines that separate $P'$ from these two
5364 integer points.
5365 $P''$ will have the same width (2) in the
5366 $x$ direction, while $P' \subset P''$.
5367 The $x$ direction is therefore also a lattice width direction of $P''$.
5368 $P''$ cannot intersect both $x=x_1$ and $x=x_1+2$ in a segment of
5369 length greater than or equal to $1$.
5370 Otherwise, it would also intersect $x=x_1+1$ in a segment of length
5371 greater than or equal to $1$.
5373 We may therefore assume that the length of the intersection
5374 of $P''$ with $x=x_1$ is smaller than $1$.
5375 If this line segment contains an integer point, then call it $y_2$.
5376 Otherwise, let $y_2$ be the greatest integer smaller than the
5377 points in the line segment.
5378 We may assume that $y_1 = y_2$.
5379 Otherwise, we can apply the unimodular transformation
5381 \begin{bmatrix}
5382 x \\
5384 \end{bmatrix}
5386 \begin{bmatrix}
5387 1 & 0 \\
5388 y_1 - y_2 & 1
5389 \end{bmatrix}
5390 \begin{bmatrix}
5391 x \\
5393 \end{bmatrix}
5396 without changing the width in direction $x$.
5397 If $P''$ contains $(x_1, y_1)$, it intersects $x=x_1$
5398 in a segment $[y_1-\alpha_1, y_1+\alpha_2]$.
5399 We may then similarly assume that $\alpha_2 \ge \alpha_1$.
5400 $P''$ will only cut $x=x_1+2$ in points with $y$-coordinate
5401 smaller than $2-\alpha_2$. The width in the $y$ direction
5402 will therefore be smaller than $2-\alpha_2+\alpha_1 \le 2$,
5403 contradicting that $x$ is a lattice width direction.
5404 If $P''$ does not contain $(x_1, y_1)$, then it only
5405 intersects $x=x_1$ in points with $y$-coordinate $y_1+\alpha$
5406 with $0 < \alpha < 1$. Given any such point, it is clear
5407 that $P''$ intersects $x=x_1+2$ only in points with $y$-coordinate
5408 strictly between $y_1-\alpha$ and $y_1+1-\alpha$, again
5409 showing that the width in the $y$ direction is smaller than $2$ and
5410 leading to the same contradiction.
5411 The contradiction shows that there can be no gaps
5412 of width 2 in the lattice width direction of $P$.
5413 \end{proof}
5414 Note that the $\omega(2)$ bound is too coarse to reach
5415 the above conclusion as $\omega(2) > 2$.
5416 An example of a polygon with lattice with greater than $2$ is the polygon
5417 with vertices $(-17/110,83/110)$, $(2/10,-9/10)$ and $(177/90, 100/90)$,
5418 shown in \autoref{f:empty:width:2}, which has width $221/110$.
5420 \begin{figure}
5421 \intercol=3cm
5422 \begin{xy}
5423 <\intercol,0pt>:<0pt,\intercol>::
5424 \def\latticebody{\POS="c"+(0,-1.5)\ar@{--}"c"+(0,1.5)}%
5425 \POS0,{\xylattice{-1}{2}00}%
5426 \def\latticebody{\POS="c"+(-1.5,0)\ar@{--}"c"+(2.5,0)}%
5427 \POS0,{\xylattice00{-1}1}%
5428 \POS@i@={(-0.1545,0.7545),(0.2,-0.9),(1.966,1.111),(-0.1545,0.74545)},{0*\xypolyline{}}
5429 \end{xy}
5430 \caption{Lattice point free polygon with lattice width 2}
5431 \label{f:empty:width:2}
5432 \end{figure}
5434 The idea of the projection algorithm
5435 is now to first find a direction in which the gaps
5436 are expected to be small and to unimodularly transform
5437 the existentially quantified variables such that this direction
5438 lies in the direction of one of the transformed variables.
5439 Then, the remaining existentially quantified variables are
5440 projected out by applying the technique recursively.
5441 The resulting generating function will have gaps at most
5442 $\omega(m)$ wide and so we have to subtract at most
5443 $\omega(m)$ shifted copies of this generating function
5444 before we can plug in 1 to project out the selected
5445 (and now only remaining) existentially quantified variable.
5446 We now look at each of these step in a bit more detail.
5448 We are given a polyhedron $P$ such that $P_{\vec s}$ is a polytope
5449 and we want to compute a generating function $f(\vec x, \vec y)$
5450 for the set $T$ in~\eqref{eq:projection:T}.
5451 We first compute the lattice width directions of
5452 the $m$-dimensional parametric polytope $P_{\vec s, \vec t}$
5453 as in \autoref{s:width}.
5454 The result is a partition of the parameter domain, i.e.,
5455 the projection of $P$ onto the first $n+d$ coordinates,
5456 into partially open polyhedra $Q_i$, together with
5457 the lattice width direction $\vec c_i$ corresponding to each $Q_i$.
5458 Since the generating functions only encode integer points,
5459 we can replace each open facet $\sp a x + b > 0$ by the closed
5460 facet $\sp a x + b - 1 \ge 0$ to obtain a collection of closed
5461 polyhedra $\tilde Q_i$. Now let
5463 P_i = P \cap \tilde Q_i \times \QQ^m
5465 and let $f_i(\vec x, \vec y)$ be the generating function of the set
5467 T_i =
5469 (\vec s, \vec t) \mid
5470 \exists \vec u \in \ZZ^m : (\vec s, \vec t, \vec u) \in P_i
5474 Then clearly,
5476 f(\vec x, \vec y) = \sum_i f_i(\vec x, \vec y)
5479 From now on, we will consider a particular $P_i$ with corresponding
5480 lattice width $\vec c_i$ and drop the $i$ subscript.
5482 We are now given a polyhedron $P$ such that the lattice width
5483 direction of $P_{\vec s, \vec t}$ is $\vec c$.
5484 We first extend $\vec c$ to an $m \times m$ unimodular matrix $U$
5485 using the technique of \autoref{s:completion},
5489 \begin{bmatrix}
5490 \vec c^\T
5493 \end{bmatrix}
5495 and then compute
5497 P' =
5498 \begin{bmatrix}
5499 I_n & 0 & 0 \\
5500 0 & I_d & 0 \\
5501 0 & 0 & U
5502 \end{bmatrix}
5506 We have
5510 (\vec s, \vec t) \mid
5511 \exists \vec u' \in \ZZ^m : (\vec s, \vec t, \vec u') \in P'
5515 i.e., we may have changed the values of the existentially
5516 quantified variables, but we have not changed the set $T$.
5517 Now consider the set
5519 T' =
5521 (\vec s, \vec t, u_1') \mid
5522 \exists (u_2',\ldots,u_m') \in \ZZ^{m-1} :
5523 (\vec s, \vec t, \vec u') \in P'
5527 This set has only $m-1$ existentially quantified variables, so
5528 we may apply this projection algorithm recursively and obtain
5529 the generating function $f'(\vec x, \vec y, z)$ for $T'$.
5530 The set $T'$ may no longer correspond to the integer points
5531 in a polytope, but, by construction, the gaps in the final
5532 coordinate are small ($\le \omega(m)$).
5534 By now we have a generating function $f'(\vec x, \vec y, z)$
5535 for the set $T'$ (with small
5536 gaps in the final coordinate) and we have to compute the
5537 generating function $f(\vec x, \vec y)$ for $T$.
5538 By computing
5539 \begin{equation}
5540 \label{eq:projection:omega}
5541 f''(\vec x, \vec y, z) =
5542 f'(\vec x, \vec y, z) \bigoplus_{k=1}^{\floor{\omega(m)}}
5543 \left( z^k f'(\vec x, \vec y, z) \right)
5545 \end{equation}
5546 where $\oplus$ represents the operation on generating functions
5547 that corresponds to set difference on the corresponding sets,
5548 we obtain a generating for the set $T''$ where only
5549 the smallest value of $u_1'$ is retained.
5550 The total number of $u_1'$s associated to any $(\vec s, \vec t)$
5551 is therefore either zero or one and so the ``multiset'' defined
5552 by taking as many copies of $(\vec s, \vec t)$ as there are
5553 associated values of $u_1'$ is actually the set $T$.
5554 That is
5556 f(\vec x, \vec y) = f''(\vec x, \vec y, 1)
5560 The only remaining problem is that the ``$\oplus$'' operation
5561 in~\eqref{eq:projection:omega} is fairly expensive.
5562 In particular, this operation is performed by first
5563 computing the \ai{Hadamard product} of the two generating functions
5564 (which corresponds to the intersection of the sets) and
5565 then subtracting the resulting generating function from this
5566 first generating function.
5567 The last operation is fairly cheap, but the Hadamard product
5568 has a time complexity which while polynomial if the dimension (in
5569 this case the maximum of $k_i$ in~\eqref{eq:rgf}) is fixed,
5570 is exponential in this dimension.
5571 Furthermore, this dimension increases by $\max k_i - d$ on each
5572 successive application of the Hadamard product, while $\max k_i > d$
5573 as soon as some projection is performed, which certainly happens in the
5574 recursive application of the algorithm.
5575 Now, the total number of Hadamard products is bounded by a constant
5576 $\floor{\omega(m)}$ and so the increase in dimension is also bounded
5577 by a constant, so the whole operation is still polynomial for
5578 fixed dimension.
5579 Nevertheless, we do not want to perform any additional Hadamard
5580 products if we do not really have to.
5581 That is, we would like to be able to detect when we can stop
5582 performing these operations {\em before} reaching the upper
5583 bound $\floor{\omega(m)}$.
5585 Let $f'_0(\vec x, \vec y, z) = f'(\vec x, \vec y, z)$ and
5586 let $f'_k(\vec x, \vec y, z)$ be the result of applying
5587 the ``set difference'' in~\eqref{eq:projection:omega} $k$ times.
5588 Denote the corresponding sets by $T'_0$ and $T'_k$.
5589 We want to find the smallest $k$ such that
5590 $f''(\vec x, \vec y, z) = f'_k(\vec x, \vec y, z)$.
5591 Note that we are talking about equality of rational functions here.
5592 Any further application of the set difference operation will
5593 not change this rational function, but it {\em will\/} typically
5594 produce a different (more complex) representation.
5595 To check whether the current $k$ is sufficient, we are going to
5596 count how many times any element of $T'_k$ still appears in a
5597 shifted copy of $T'_0$, with shift greater than or equal to $k+1$.
5598 If this number is zero, any further set difference will have no effect.
5599 That is, we want to compute
5601 \sum_{l=k+1}^\infty
5602 \left|
5603 T'_l \cap \left(\vec e_{n+d+1} + T' \right)
5604 \right|
5607 or, in terms of generating functions,
5609 h(\vec x, \vec y, z) = \sum_{l=k+1}^\infty
5610 f_k'(\vec x, \vec y, z) \star z^l \, f'(\vec x, \vec y, z)
5613 We should point out here that while the Hadamard product ($\star$)
5614 corresponds to intersection when applied to generator functions
5615 of indicator functions (i.e., with coefficients one or zero),
5616 in general it will produce a generating function with coefficients
5617 that are the product of the corresponding coefficients in the two
5618 operands.
5619 We can therefore rewrite the above equation as
5621 \begin{aligned}
5622 h(\vec x, \vec y, z) & = \sum_{l=k+1}^\infty
5623 f_k'(\vec x, \vec y, z) \star z^l \, f'(\vec x, \vec y, z)
5625 & = f_k'(\vec x, \vec y, z) \star
5626 \left(
5627 \sum_{l=k+1}^\infty z^l \, f'(\vec x, \vec y, z)
5628 \right)
5630 & = f_k'(\vec x, \vec y, z) \star
5631 \frac{z^{k+1} \, f'(\vec x, \vec y, z)}{1-z}
5633 \end{aligned}
5635 Computing $h(\vec x, \vec y, 1)$ would give us a generating
5636 function with as coefficients how many times a point of $T'_k$
5637 still appears in a shifted copy of $T'_0$ for any particular
5638 value of $\vec s$ and $\vec t$.
5639 However, we want to know if this number is zero for {\em all\/}
5640 values of $\vec s$ and $\vec t$, so we compute $h(\vec 1, \vec 1, 1)$
5641 instead. We have to be careful here since we allow the polyhedron
5642 $P$ to be unbounded and so we should apply the technique
5643 of \autoref{s:infinite} with $Q$ the projection of $P$ on
5644 the remaining coordinates.
5646 Note that testing whether we can stop is more expensive
5647 than applying the next iteration (since we have an extra
5648 $(1-z)$ factor in the denominator of one of the operands).
5649 However, we may save many iterations by stopping early
5650 and we will not needlessly replace a given representation
5651 of $f''(\vec x, \vec y, z)$ by a more complex representation
5652 (with more factors in the denominator).
5653 An alternative way of checking whether we can stop is to
5654 test whether $f'_k(\vec x, \vec y, z) = f'_{k+1}(\vec x, \vec y, z)$
5655 (as rational functions).
5656 To do so, we would need to check that both
5658 f'_k(\vec x, \vec y, z) -
5659 \left( f'_k(\vec x, \vec y, z) \star f'_{k+1}(\vec x, \vec y, z) \right)
5663 f'_{k+1}(\vec x, \vec y, z) -
5664 \left( f'_k(\vec x, \vec y, z) \star f'_{k+1}(\vec x, \vec y, z) \right)
5666 are zero and this Hadamard product is more expensive than
5667 the one above.
5669 \begin{figure}
5670 \intercol=1.05cm
5671 \begin{xy}
5672 <\intercol,0pt>:<0pt,\intercol>::
5673 \def\latticebody{\POS="c"+(0,-5.5)\ar@{--}"c"+(0,5.5)}%
5674 \POS0,{\xylattice{-5}{5}00}%
5675 \def\latticebody{\POS="c"+(-5.5,0)\ar@{--}"c"+(5.5,0)}%
5676 \POS0,{\xylattice00{-5}5}%
5677 \POS(0,-5.5)\ar(0,5.5) \POS(0,5.5)*+!UL{x_2}
5678 \POS(-5.5,0)\ar(5.5,0) \POS(5.5,0)*+!DR{x_1}
5679 \POS@i@={(-5,0),(5,0)},{0*[|(2)]\xypolyline{}}
5680 \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{}}
5681 \POS@i@={(-4,1),(4,1),(4,-1),(-4,-1),(-4,1)},{0*[|(2)]\xypolyline{}}
5682 \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{}}
5683 \POS@i@={(-3,2),(3,2),(3,-2),(-3,-2),(-3,2)},{0*[|(2)]\xypolyline{}}
5684 \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{}}
5685 \POS@i@={(-2,3),(2,3),(2,-3),(-2,-3),(-2,3)},{0*[|(2)]\xypolyline{}}
5686 \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{}}
5687 \POS@i@={(-1,4),(1,4),(1,-4),(-1,-4),(-1,4)},{0*[|(2)]\xypolyline{}}
5688 \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{}}
5689 \POS@i@={(0,-5),(0,5)},{0*[|(2)]\xypolyline{}}
5690 \POS(-5,0)*+!DR{5}
5691 \POS(-4.5,0.5)*+!DR{4}
5692 \POS(-4,1)*+!DR{3}
5693 \POS(-3.5,1.5)*+!DR{2}
5694 \POS(-3,2)*+!DR{1}
5695 \POS(-2.5,2.5)*+!DR{0}
5696 \POS(-2,3)*+!DR{-1}
5697 \POS(-1.5,3.5)*+!DR{-2}
5698 \POS(-1,4)*+!DR{-3}
5699 \POS(-0.5,4.5)*+!DR{-4}
5700 \POS(-0,5)*+!DR{-5}
5701 \end{xy}
5702 \caption{The parametric polytope from Example~\ref{ex:projection}
5703 for different values of the parameter}
5704 \label{f:ex:projection}
5705 \end{figure}
5707 \begin{example} \label{ex:projection}
5708 Consider once more the parametric polytope
5710 P(p) = \left\{\,
5711 \vec x \mid
5712 \begin{aligned}
5713 -2 x_1 + p + 5 &\ge 0 \\
5714 2 x_1 + p + 5 &\ge 0 \\
5715 -2 x_2 - p + 5 &\ge 0 \\
5716 2 x_2 - p + 5 &\ge 0
5717 \end{aligned}
5718 \,\right\}
5720 from \shortciteN[Example~2.1.7]{Woods2004PhD}
5721 and Example~\ref{ex:width:2} and assume we want to
5722 compute
5724 c(p) = \left[ \exists \vec x \in \ZZ^2: (p, \vec x) \in P \right]
5727 That is, we simply want to know for which values of $p$
5728 the polytope is non-empty.
5729 Now, there are more efficient ways of answering this particular question,
5730 but we will use it here as an example of the above algorithm.
5731 The polytope $P(p)$ is shown in \autoref{f:ex:projection} for
5732 all integer value of the parameter for which the polytope
5733 is non-empty.
5735 \begin{figure}
5736 \intercol=1.05cm
5737 \begin{xy}
5738 <\intercol,0pt>:<0pt,\intercol>::
5739 \def\latticebody{\POS="c"+(0,-5.5)\ar@{--}"c"+(0,5.5)}%
5740 \POS0,{\xylattice{-5}{5}00}%
5741 \def\latticebody{\POS="c"+(-5.5,0)\ar@{--}"c"+(5.5,0)}%
5742 \POS0,{\xylattice00{-5}5}%
5743 \POS(0,-5.5)\ar(0,5.5) \POS(0,5.5)*+!UL{x_2}
5744 \POS(-5.5,0)\ar(5.5,0) \POS(5.5,0)*+!DR{x_1}
5745 \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{}}
5746 \POS@i@={(-2,3),(2,3),(2,-3),(-2,-3),(-2,3)},{0*[|(2)]\xypolyline{}}
5747 \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{}}
5748 \POS@i@={(-1,4),(1,4),(1,-4),(-1,-4),(-1,4)},{0*[|(2)]\xypolyline{}}
5749 \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{}}
5750 \POS@i@={(0,-5),(0,5)},{0*[|(2)]\xypolyline{}}
5751 \POS(-2.5,2.5)*+!DR{0}
5752 \POS(-2,3)*+!DR{1}
5753 \POS(-1.5,3.5)*+!DR{2}
5754 \POS(-1,4)*+!DR{3}
5755 \POS(-0.5,4.5)*+!DR{4}
5756 \POS(-0,5)*+!DR{5}
5757 \POS(0,-11.5)*\xybox{
5758 \def\latticebody{\POS="c"+(0,-0.5)\ar@{--}"c"+(0,5.5)}%
5759 \POS0,{\xylattice{-5}{5}00}%
5760 \def\latticebody{\POS="c"+(-5.5,0)\ar@{--}"c"+(5.5,0)}%
5761 \POS0,{\xylattice00{0}5}%
5762 \POS(0,-0.5)\ar(0,5.5) \POS(0,5.5)*+!UR{p}
5763 \POS(-5.5,0)\ar(5.5,0) \POS(5.5,0)*+!UR{x_1}
5764 \POS(-2,0)*{\bullet}
5765 \POS(-1,0)*{\bullet},*+!DL{1}
5766 \POS(-0,0)*{\bullet},*+!DL{2}
5767 \POS(1,0)*{\bullet},*+!DL{3}
5768 \POS(2,0)*{\bullet},*+!DL{4}
5769 \POS(3,0)*+!DL{5}
5770 \POS(4,0)*+!DL{5}
5771 \POS(5,0)*+!DL{5}
5772 \POS(-2,1)*{\bullet}
5773 \POS(-1,1)*{\bullet},*+!DL{1}
5774 \POS(-0,1)*{\bullet},*+!DL{2}
5775 \POS(1,1)*{\bullet},*+!DL{3}
5776 \POS(2,1)*{\bullet},*+!DL{4}
5777 \POS(3,1)*+!DL{5}
5778 \POS(4,1)*+!DL{5}
5779 \POS(5,1)*+!DL{5}
5780 \POS(-1,2)*{\bullet}
5781 \POS(-0,2)*{\bullet},*+!DL{1}
5782 \POS(1,2)*{\bullet},*+!DL{2}
5783 \POS(2,2)*+!DL{3}
5784 \POS(3,2)*+!DL{3}
5785 \POS(4,2)*+!DL{3}
5786 \POS(5,2)*+!DL{3}
5787 \POS(-1,3)*{\bullet}
5788 \POS(-0,3)*{\bullet},*+!DL{1}
5789 \POS(1,3)*{\bullet},*+!DL{2}
5790 \POS(2,3)*+!DL{3}
5791 \POS(3,3)*+!DL{3}
5792 \POS(4,3)*+!DL{3}
5793 \POS(5,3)*+!DL{3}
5794 \POS(-0,4)*{\bullet}
5795 \POS(1,4)*+!DL{1}
5796 \POS(2,4)*+!DL{1}
5797 \POS(3,4)*+!DL{1}
5798 \POS(4,4)*+!DL{1}
5799 \POS(5,4)*+!DL{1}
5800 \POS(-0,5)*{\bullet}
5801 \POS(1,5)*+!DL{1}
5802 \POS(2,5)*+!DL{1}
5803 \POS(3,5)*+!DL{1}
5804 \POS(4,5)*+!DL{1}
5805 \POS(5,5)*+!DL{1}
5806 \POS(-6,-0.5)\ar(-6,5.5) \POS(-6,5.5)*+!UL{p}
5807 \POS(-6,0)*{\bullet}
5808 \POS(-6,1)*{\bullet}
5809 \POS(-6,2)*{\bullet}
5810 \POS(-6,3)*{\bullet}
5811 \POS(-6,4)*{\bullet}
5812 \POS(-6,5)*{\bullet}
5814 \end{xy}
5815 \caption{The transformed parametric polytope from Example~\ref{ex:projection}
5816 for $0 \le p \le 5$}
5817 \label{f:ex:projection:transformed}
5818 \end{figure}
5820 The first step is to compute the lattice width of $P(p)$.
5821 In Example~\ref{ex:width:2}, we already obtained the decomposition
5822 of the parameter domain into
5824 \begin{aligned}
5825 \hat C_1 &= \{\, p \mid 0 \le p \le 5 \,\} \\
5826 \hat C_2 &= \{\, p \mid -5 \le p \le -1 \,\}
5828 \end{aligned}
5830 with corresponding lattice widths and lattice width directions
5832 \begin{aligned}
5833 \vec c_1 &= (0,1) & w_{\vec c_1}(p) &= 5-p \\
5834 \vec c_2 &= (1,0) & w_{\vec c_2}(p) &= 5+p
5836 \end{aligned}
5838 Note that in this example, the gaps in both coordinate directions
5839 are $1$, so, in principle, there is no need to perform any unimodular
5840 transformation here. Nevertheless, we will apply the transformation
5841 that would be applied by the algorithm.
5842 On the first domain, we extend the lattice width direction
5843 to the unimodular matrix
5845 \begin{bmatrix}
5846 0 & 1 \\
5847 1 & 0
5848 \end{bmatrix}
5851 After application to the existentially quantified variables $\vec x$,
5852 we obtain the parametric polytope
5854 P'_1(p) = \left\{\,
5855 \vec x \mid
5856 \begin{aligned}
5857 -2 x_2 + p + 5 &\ge 0 \\
5858 2 x_2 + p + 5 &\ge 0 \\
5859 -2 x_1 - p + 5 &\ge 0 \\
5860 2 x_1 - p + 5 &\ge 0 \\
5861 p & \ge 0 \\
5862 \end{aligned}
5863 \,\right\}
5865 shown in the top part of \autoref{f:ex:projection:transformed}.
5866 We now temporarily remove the existential quantification on $x_1$,
5867 resulting in
5869 T' = \{ (p, x_1) \in \ZZ^2 \mid \exists x_2 \in \ZZ : (s, \vec x) \in P' \}
5872 Since there is only one existentially quantified variable left,
5873 we know we only have to shift and subtract the set once to obtain
5874 a set with at most one value of $x_2$ associated to each value
5875 of $(p, x_1)$.
5876 In particular, let $f(x,z_1,z_2)$ be the generating function
5877 of the integer points in $P'$. Then $g(x,z_1) = f'(x,z_1,1)$, with
5878 $f'(x,z_1,z_2) = f(x,z_1,z_2) - f(x,z_1,z_2) \star z_2 f(x,z_1,z_2)$,
5879 is the generating function of $T'$.
5881 To check whether we need to subtract any shifted copies of
5882 $g(x,z_1)$ from itself, we compute
5884 h(x,z_1) = g(x,z_1) \star \frac{z_1 \, g(x,z_1)}{1-z_1}
5887 The second argument of this Hadamard product is depicted
5888 in \autoref{f:ex:projection:transformed} by its coefficients.
5889 The exponents in $h(x,z_1)$ that have a non-zero coefficient
5890 are therefore those marked by both a dot ($\bullet$) and
5891 a number. The total sum can be computed as $h(1,1) = 26$,
5892 which is finite, but non-zero. We therefore need to subtract
5893 at least one shifted copy of $g(x,z_1)$.
5896 g'(x,z_1) = g(x,z_1) - g(x,z_1) \star z_1 g(x,z_1)
5899 Computing
5901 h'(x,z_1) = g'(x,z_1) \star \frac{z_1^2 \, g(x,z_1)}{1-z_1}
5904 we would find that $h'(1,1) = 0$ and so we do not need
5905 to shift and subtract any further.
5906 However, since we are dealing with a two-dimensional problem,
5907 we already know from \autoref{l:gap} that we can stop
5908 after one shift and subtract, so we do not even need
5909 to compute $h'(x,z_1)$ here.
5910 The function $g'(x,z_1)$ now only has non-zero coefficients
5911 for at most one exponent of $z_1$ for each exponent of $x$.
5912 We may therefore project down to obtain
5913 the function $g'(x,1)$, which is the generating function
5914 of the set in the lower left part of \autoref{f:ex:projection:transformed}.
5916 For the chamber $\hat C_2$ of the parameter domain, the computations
5917 are nearly identical and the final result is simply the sum
5918 of the two generating functions found for each of the two (disjoint)
5919 chambers.
5921 \end{example}