1: \NeedsTeXFormat{LaTeX2e}[1994/06/01]
2: \documentclass[11pt,a4paper]{article}
3:
4: %\RequirePackage[T1]{fontenc}
5: %\RequirePackage[latin2]{inputenc}
6: %\RequirePackage{times}
7: \usepackage{latexsym}
8: \usepackage{amssymb}
9: \usepackage{amsmath}
10: \usepackage{amsthm}
11: \usepackage{eucal}
12: \usepackage{bezier}
13: \usepackage[matrix,curve,arrow,tips,frame]{xy}
14:
15:
16: \theoremstyle{plain} \newtheorem{theorem}{Theorem}
17: \newtheorem{lemma}[theorem]{Lemma}
18: \newtheorem{sublemma}[theorem]{Sublemma}
19: \newtheorem{proposition}[theorem]{Proposition}
20: \newtheorem{corollary}[theorem]{Corollary} \newtheorem{cl}{Claim}
21: \newtheorem{conj}{Conjecture} \newtheorem*{observation}{Observation}
22:
23: \theoremstyle{definition} \newtheorem{definition}[theorem]{Definition}
24: \newtheorem{algorithm}{Algorithm}
25: \renewcommand{\thealgorithm}{\Alph{algorithm}}
26:
27: \theoremstyle{remark} \newtheorem*{claim}{Claim}
28: \newtheorem*{example}{Example}
29: \newtheorem*{acknowledgement}{Acknowledgement}
30:
31: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32: %
33: % macros for structures, models, algebras...
34: %
35: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36:
37:
38: \newcommand{\N}{\mathbb{N}}
39: \newcommand{\Z}{\mathbb{Z}}
40:
41: %\newcommand{\C}{\mathbb{C}}
42: %\DeclareMathOperator{\pr}{\mathrm{pr}}
43:
44: \newcommand{\yes}{\mathbf{yes}}
45: \newcommand{\no}{\mathbf{no}}
46:
47: \newcommand{\nil}{\mathbf{nil}}
48: \newcommand{\answ}{\mathbf{answ}}
49: \DeclareMathOperator{\PP}{\mathcal{P}}
50: \DeclareMathOperator{\LL}{\mathcal{L}}
51: \DeclareMathOperator{\intersect}{\cap}
52:
53: \newcommand{\NP}{\mathcal{NP}}
54: \newcommand{\coNP}{\mathrm{co}\mathcal{NP}}
55: \newcommand{\EXPTIME}{\mathcal{EXPTIME}}
56: \newcommand{\PTIME}{\mathcal{PTIME}}
57: \newcommand{\PSPACE}{\mathcal{PSPACE}}
58:
59:
60: \newcommand{\X}{\mathcal{X}}
61: \newcommand{\Y}{\mathcal{Y}}
62:
63:
64: \newcommand{\Lor}{\bigvee}
65: \newcommand{\Land}{\bigwedge}
66:
67: \DeclareMathOperator{\U}{\mathsf{Until}}
68: \DeclareMathOperator{\Since}{\mathsf{Since}}
69: \DeclareMathOperator{\NEXT}{\bigcirc}
70: \DeclareMathOperator{\FDIA}{\Diamond}
71: \DeclareMathOperator{\FBOX}{\Box}
72: \DeclareMathOperator{\PREV}{\unitlength 1.00pt\begin{picture}(8.00,10.00)\put(4.00,3.00){\circle*{6}}\end{picture}}
73: \DeclareMathOperator{\PDIA}{\blacklozenge}
74: \DeclareMathOperator{\PBOX}{\blacksquare}
75:
76: \newcommand{\TL}{\mathrm{TL}}
77: \newcommand{\M}{\mathcal{M}}
78:
79: \newcommand{\conn}[9]{
80: \begin{array}{|c|ccc|} \hline
81: \multicolumn{4}{|c|}{x #1 y}\\ \hline\hline x\diagdown y &0 &1 &\bot\\\hline
82: 0   \\ 1   \\ \bot  	
83: &\bot\\ \hline
84: \end{array}}
85:
86: \newcommand{\varconn}[9]{
87: \begin{array}{|c|c|c|c|} \hline
88: \multicolumn{4}{|c|}{#1(x,y)}\\ \hline\hline x\diagdown y &0 &1 &\bot\\
89: \hline 0   \\ \hline 1   \\ \hline \bot  	
90: &\bot\\ \hline
91: \end{array}}
92:
93: \newcommand{\SAC}{\mathrm{SAC}}
94: \newcommand{\GNW}{\mathrm{GNW}}
95: \newcommand{\PS}{\mathrm{PS}}
96: \newcommand{\CL}{\mathrm{CL}}
97: \newcommand{\Sch}{\mathrm{Sch}}
98: \newcommand{\Cal}{\mathrm{Cal}}
99:
100: \newcommand{\lra}{\leftrightarrow}
101:
102: \newcommand{\true}{1}
103: \newcommand{\false}{0}
104: \renewcommand{\Cup}{\bigcup}
105: \renewcommand{\Cap}{\bigcap}
106: \DeclareMathOperator{\falka}{\leftrightsquigarrow}
107:
108: \newcommand{\tm}{${}^{\mathrm{TM}}$}
109: \newcommand{\A}{\mathfrak{A}}
110: \newcommand{\B}{\mathfrak{B}}
111: \newcommand{\E}{\mathcal{E}}
112:
113: \newcommand{\cea}{{\em cea\/}}
114: \newcommand{\fM}{\mathfrak{M}}
115: \newcommand{\RR}{{\mathsf{R}}}
116: \newcommand{\RS}{{\mathsf{RS}}}
117: \renewcommand{\S}{{\mathsf{S}}}
118: \newcommand{\R}{{\mathsf{R}}}%{{\scalebox{.55}[.77]{$\mathsf{R}$}}}
119: \newcommand{\C}{\complement}
120: \newcommand{\trzy}{{\text{\boldmath$\mathsf{3}$}}}
121: \newcommand{\dwa}{{\text{\boldmath$\mathsf{2}$}}}
122: \DeclareMathOperator{\hatpr}{{\underline{Prob}}}
123: \newcommand{\hatOmega}{\underline{\Omega}}
124: \newcommand{\PC}{\mathcal{PC}}
125: \newcommand{\CC}{\mathcal{C}}
126: \renewcommand{\P}{\mathcal{P}}
127: \newcommand{\lland}{\curlywedge}
128: \newcommand{\llor}{\curlyvee}
129: \newcommand{\LLor}{\bigvee}%{\scalebox{1.5}{$\llor$}}
130: \DeclareMathOperator{\llto}{\rightarrowtail}
131: \newcommand{\first}{\mathrm{first}}
132: \newcommand{\last}{\mathrm{last}}
133: \newcommand{\ce}[1]{[\![#1]\!]}
134: \newcommand{\ps}[1]{\langle\!\langle#1\rangle\!\rangle_\PS}
135: \newcommand{\gnw}[1]{\langle\!\langle#1\rangle\!\rangle_\GNW}
136: \newcommand{\sac}[1]{\langle\!\langle#1\rangle\!\rangle_\SAC}
137: \renewcommand{\cl}[1]{\langle\!\langle#1\rangle\!\rangle_\CL}
138: \newcommand{\ttt}[1]{\tau(#1)}
139: \newcommand{\ttl}[1]{\sigma(#1)}
140: \newcommand{\ttlr}[1]{\sigma^{\R}(#1)}
141: \newcommand{\ttr}[1]{\tau^{\RR}(#1)}
142: \newcommand{\ttlrs}[1]{\sigma^{\RS}(#1)}
143: \newcommand{\ttrs}[1]{\tau^{\RS}(#1)}
144: \newcommand{\TT}{{(\TL|\TL)}}
145: \newcommand{\pr}{\Pr\nolimits}
146: \newcommand{\defn}{\mathop{\uparrow}}
147: \newcommand{\TRUE}{\mathit{true}}
148: \newcommand{\FALSE}{\mathit{false}}
149: \newcommand{\atone}{\mathbf{@}_1}
150: \newcommand{\attwo}{\mathbf{@}_2}
151: \newcommand{\llnot}{\sim\!}
152: \parindent0pt
153: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
154: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
155: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156:
157:
158: %\usepackage{color}
159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
162: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
163: \title{Embedding Conditional Event Algebras Into Temporal Calculus
164: of Conditionals}
165: \author{
166: Jerzy Tyszkiewicz$^{1,2}$\\
167: Achim Hoffmann$^2$\\
168: Arthur Ramer$^2$}
169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
170: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
172: %\CompileMatrices
173: \pagestyle{myheadings}
174: %\markright{{\tt \jobname .tex}, version of \today}
175: \begin{document}
176: \begin{titlepage}
177: \maketitle
178: \begin{center}$^1$ Institute of Informatics,\\ Warsaw University,\\
179: Banacha 2,\\ 02-097 Warszawa,\\ Poland.\\ E-mail {\tt
180: jty@mimuw.edu.pl}.\\ Supported by the Polish Research Council KBN grant
181: 8 T11C 027 16.\\[5pt] $^2$ School CSE,\\ UNSW,\\ 2052 Sydney,\\
182: Australia.\\ E-mail {\tt \{jty|achim|ramer\}@cse.unsw.edu.au}.\\
183: Supported by the Australian Research Council ARC grant A 49800112
184: (1998--2000).
185: \end{center}
186: \thispagestyle{empty}
187: \end{titlepage}
188: \begin{titlepage}
189: \begin{abstract} In this paper we prove that all the existing {\em
190: conditional event algebras\/} (abbreviated \cea{} in this paper) embed
191: into the three-valued extension $\TT$ of temporal logic of discrete past
192: time, which the authors of this paper have proposed in \cite{TT} as a
193: general model of conditional events.
194:
195: First of all, we discuss the descriptive incompleteness of the cea's.
196: In this direction, we show that some important notions, like
197: independence of conditional events, cannot be properly addressed in
198: the \cea{} framework, while they can be precisely formulated and
199: analyzed in the $\TT$ setting.
200:
201: We also demonstrate that the embeddings allow one to use the native
202: $\TT$ algorithms for computing probabilities of complex conditional
203: expressions of the embedded \cea's, and that these algorithms can
204: outperform those previously known.
205: \end{abstract}
206: \end{titlepage}
207:
208: \begin{titlepage}
209: \tableofcontents
210: \listoffigures
211: \end{titlepage}
212:
213: \section{Preliminaries and statement of the problem}
214:
215: \subsection{The problem of conditional objects}
216:
217: Probabilistic reasoning \cite{p88} is the basis of Bayesian methods of
218: expert system inferences, of knowledge discovery in databases, and in
219: several other domains of computer, information, and decision
220: sciences. The model of conditioning and conditional objects we discuss
221: serves equally to reason about probabilities over a finite domain $X$,
222: or probabilistic propositional logic with a finite set of atomic
223: formulae.
224:
225:
226: Computing of conditional probabilities of the form
227: $\Pr(X|Y_1,\dots,Y_n)$ and, by extension of conditional beliefs, is well
228: understood. Attempts of defining first the {\em conditional objects\/}
229: of the basic form $X|Y$, and then defining $\Pr(X|Y)$ as $\Pr((X|Y))$
230: were proposed, without much success, by some of the founders of
231: probability \cite{b57,d72}. They were taken up systematically only
232: about 1980 \cite{a86,l76,es94,g91,gn96}. The development was slow, both
233: because of logical difficulties --- interpretation of conditionals, and
234: even more because the computational model is difficult to construct.
235: (While $a|b$ appears to stand for a sentence `if $b$ then $a$' and the
236: probability is $\Pr(a|b)=\Pr(a\land b)/\Pr(b)$, there is no obvious
237: calculation for $\Pr(a|(b|c))$, nor intuitive meaning for $a|(b|c),$
238: $(a|b)\land(c|d),$ and the like.)
239:
240: The idea of defining conditional objects was entertained by some
241: founders of modern probability \cite{b57,d72}, but generally abandoned
242: since introduction of the measure-theoretic model. It was revived
243: mostly by philosophers in 1970's \cite{a86,v77} with a view towards
244: artificial intelligence reasoning. Formal computational models came in
245: the late 1980's \cite{c87,gnw91} with only one, based on formal
246: fractions and three-valued indicator functions, used for few actual
247: calculations of conditionals and their probabilities. That model may
248: give results whose values are open to questions \cite{c94}.
249:
250: The authors of this paper have developed in the companion paper
251: \cite{TT} a temporal calculus $\TT$ of conditionals, based on
252: the early ideas of de Finetti \cite{d72}.
253:
254: In the present paper we show that all the major previously existing
255: systems of conditionals, the so called {\em conditional event algebras}
256: (see \cite{kniga} or Sect.\ \ref{cea} in the current paper), embed
257: isomorphically into $\TT.$ Looking at them as fragments of $\TT,$ we
258: demonstrate their insufficient expressive power and other defects in
259: their construction.
260:
261: They attempt to provide certain kind of a model of the logic of
262: conditional expressions, built up from simple conditionals of the form
263: $(a|b)$ with the connectives: conjunction, disjunction and
264: complementation.
265:
266: However, the semantical objects assigned to the expressions are not
267: required to be of probabilistic nature, so they fail to provide
268: methods to verify the chosen structure experimentally.
269:
270: The structure of conditionals is not determined functorially by the
271: space of nonconditional events.
272:
273: Moreover, very restricted setting of \cea's does not allow one to
274: address many important questions, like stochastic independence of
275: complex conditionals. In the $\TT$ setting independence can be precisely
276: defined and analyzed, unlike in the \cea\ formalism, (cf.\ Theorem
277: \ref{PS-indep} below).
278:
279: Finally, we use the algorithms for calculating probabilities
280: in $\TT,$ which stem from the highly developed algorithms for
281: calculating limiting probabilities in Markov chains, and apply them to
282: the embedded \cea's. It appears that these algorithms clearly
283: outperform the previously known ones for the important {\em product
284: space} \cea{} \cite{g94}.
285:
286: Consequently, we believe that our $\TT$ can be used as a single
287: alternative to each of the major \cea's considered in the literature so
288: far, superior to each of them, in the sense of expressive power,
289: clarity of logical and semantical structure, and, last but not least,
290: availability of efficient algorithms.
291:
292: \section{The tools}
293: \subsection{$\TT$, Moore machines and Markov chains}
294:
295: We describe briefly the construction of temporal conditionals,
296: presented in detail in \cite{TT}.
297:
298: Let $\E=\{a,b,c,d,\dots\}$ be a finite set of basic events, and let
299: $\Sigma$ be the Boolean algebra generated by $\E,$ and $\Omega$ the set
300: of atoms of $\Sigma.$ Consequently, $\Sigma$ is isomorphic to the
301: powerset of $\Omega,$ and $\Omega$ is isomorphic to the powerset of
302: $\E.$ Any element of $\Sigma$ will be considered as an event, and, in
303: particular, $\E\subseteq\Sigma.$
304:
305: The union, intersection and complementation in $\Sigma$ are denoted by
306: $a\cup b,\ a\cap b$ and $a^\C,$ respectively. The least and greatest
307: elements of $\Sigma$ are denoted $\varnothing$ and $\Omega,$
308: respectively. However, sometimes we use a more compact notation,
309: replacing $\cap$ by a juxtaposition. When we turn to logic, it is
310: customary to use yet another notation: $a\lor b,\ a\land b$ and $\lnot
311: a,$ respectively. In this situation $\Omega$ appears as $\TRUE$ and
312: $\varnothing$ as $\FALSE,$ but $1$ and $0,$ respectively, are
313: incidentally used, as well.
314:
315: $\trzy =\{0,1,\bot\}$ is the set of truth values, interpreted as {\em
316: true}, {\em false\/} and {\em undefined}, respectively. The subset of
317: $\trzy $ consisting of $0$ and $1$ will be denoted $\dwa.$
318:
319: \bigskip
320:
321: Let us first define {\em temporal logic of linear discrete past time},
322: called $\TL.$
323:
324: The formulas are built up from the set $\E$ (the same set of basic
325: events as before), interpreted as propositional variables here, and
326: are closed under the following formula formation rules:
327:
328: \begin{enumerate}
329: \item Every $a\in \E$ is a formula of temporal logic.
330: \item If $\varphi,\psi\in \TL,$ then their boolean combinations
331: $\varphi\lor \psi$ $\lnot\varphi$ are in $\TL.$ The other Boolean
332: connectives: $\land,\to,\lra,\dots$ can be defined in terms of $\lnot$
333: and $\lor,$ as usual.
334:
335: \item If $\varphi,\psi\in \TL,$ then their past tense temporal
336: combinations $\PREV\varphi$ and $\varphi\Since\psi$ are in $\TL,$
337: where $\PREV\varphi$ is spelled ``previously $\varphi.$''
338: \end{enumerate}
339:
340: A model of temporal logic is a sequence $\M=s_0,s_1,\dots,s_n$ of
341: states, each state being a function from $\E$ (the same set of basic
342: events as before) to the boolean values $\{0,1\}.$ Note that a state
343: can be therefore understood as an atomic event from $\Omega,$ and $\M$
344: can be thought of as a word from $\Omega^+.$ The states of $\M$ are
345: ordered by $\leq,$ and $s+1$ denotes the successor state of $s.$ We
346: adopt the convention that, unless explicitly indicated otherwise, a
347: model is always of length $n+1,$ and thus $n$ is always the last state
348: of a model.
349:
350: For every state $s$ of $\M$ we define inductively what it means that a
351: formula $\varphi\in\TL$ is satisfied in the state $s$ of $\M,$
352: symbolically $\M,s\models\varphi.$
353:
354: \begin{enumerate}
355: \item $\M,s\models a$ iff $s(a)=1$
356: \item
357: \begin{align*}
358: \M,s\models\lnot\varphi&:\iff\ \M,s\not\models\varphi,\\
359: \M,s\models\varphi\lor\psi&:\iff
360: \M,s\models\varphi\ \text{or}\ \M,s\models\psi.
361: \end{align*}
362:
363:
364: \item
365: \begin{align*}
366: \M,s\models\PREV\varphi&:\iff s>0\ \text{and}\ \M,s-1\models\varphi;\\
367: \M,s\models\varphi\Since\psi&:\iff(\exists t\leq s)(\M,t\models\psi\
368: \text{and}\ (\forall t< w\leq s)M,w\models\varphi).
369: \end{align*}
370: \end{enumerate}
371:
372: The syntactic abbreviations $\PBOX\varphi$ and $\PDIA\varphi$ are of
373: common use in $\TL.$ They are defined by $\PDIA\varphi\equiv\TRUE
374: \Since\varphi$ and $\PBOX\varphi\equiv\lnot\PDIA\lnot\varphi.$ The
375: first of them is spelled ``once $\varphi$'' and the latter ``always in
376: the past $\varphi$''.
377:
378: Their semantics is then equivalent to
379: \begin{align*}
380: \M,s\models\PBOX\varphi&\iff(\forall t\leq s)\M,t\models\varphi;\\
381: \M,s\models\PDIA\varphi&\iff(\exists t\leq s) \M,t\models\varphi.
382: \end{align*}
383:
384: \begin{theorem}[see \cite{Emerson}]\label{Emerson}
385: The set of valid $\TL$ formulas is complete in $\PSPACE.$ The set of
386: valid $\TL$ formulas with $\PBOX$ and $\PDIA$ as the only temporal
387: connectives is complete in $\coNP.$
388: \end{theorem}
389:
390: $\TT$ is the logic of formulas of the form $(\varphi|\psi),$ where
391: $\varphi,\psi\in\TL.$ $\TT$ is a $\trzy$-valued extension of $\TL,$
392: and $(\varphi|\psi)$ is
393:
394: \begin{itemize}
395: \item[$1).$] true in $\M,n$ iff $\M,n\models\varphi\land\psi.$
396: \item[$0).$] false in $\M,n$ iff $\M,n\models\lnot\varphi\land\psi.$
397: \item[$\bot).$] undefined in $\M,n$ iff $\M,n\models\lnot\psi.$
398: \end{itemize}
399:
400: \bigskip
401:
402:
403: A {\em $\trzy$-valued Moore machine $\A$} is a five-tuple
404: $\A=(Q,\Omega,\delta,h,q_0),$ where where $Q$ is its set of states,
405: $\Omega$ (the same set of atomic events as before) is the input
406: alphabet, $q_0\in Q$ is the initial state and $\delta: Q
407: \times\Omega\to Q$ is the transition function, and $h$ is the output
408: function $Q\to\trzy.$
409:
410: Formally, to describe the computation of $\A$ we extend $\delta$ to a
411: function $\hat{\delta}: Q \times\Omega^+\to Q$ in the
412: following way:
413:
414: \[\hat{\delta}(q,w)=\begin{cases}
415: \delta(q,w)&\text{if $|w|=1$}\\
416: \delta(\hat{\delta}(q,v),\omega)&\text{if $w=v\omega.$}
417: \end{cases}
418: \]
419:
420:
421:
422: $\A$ computes a function $f_\A:\Omega^+\to \trzy^+$ defined by
423:
424: \[
425: f_\A(\omega_1\omega_2\dots\omega_n)
426: =h(\hat{\delta}(q_0,\omega_1))h(\hat{\delta}(q_0,\omega_1\omega_2))\dots
427: h(\hat{\delta}(q_0,\omega_1\omega_2\dots\omega_n))
428: \]
429:
430:
431: (note that $|f_\A(\omega_1\omega_2\dots\omega_n)|=n,$ as desired)
432:
433: We picture $\A$ as a labeled directed graph, whose vertices are
434: elements of $Q,$ labeled by their values under $h,$ the function
435: $\delta$ is represented by directed edges labeled by elements of
436: $\Omega$: the edge labeled by $\omega\in\Omega$ from $q\in Q$ leads
437: to $\delta(q,\omega).$ The initial state is typically indicated by an
438: unlabeled edge ``from nowhere'' to this state.
439:
440: As the letters of the input word $w\in\Omega^+$ come in one after
441: another, we walk in the graph, always choosing the edge labeled by
442: the letter we receive. At each step it reports to the outside world
443: the value $h(q)$ of the state $q$ in which it is at the moment.
444:
445: Drawing Moore machines, we almost always make certain graphical
446: simplification: we merge all the transitions joining the same pair of
447: states into a single transition, labeled by the union (evaluated in
448: $\Sigma$) of all the labels. Sometimes we go even farther and drop the
449: label altogether from one transition, which means that all the
450: remaining input letters follow this transition.
451:
452:
453: It is known for deterministic finite automata \cite{HU}, and extends
454: easily to Moore machines with the same proof, that for any such device
455: there is a unique (up to isomorphism) {\em minimal\/} (with respect to
456: the number of states) device of the same kind, which accepts the same
457: language (computes the same function, respectively). Moreover, this
458: minimal device can be obtained from any such device as a quotient
459: automaton/machine, i.e., by dividing the state space by some equivalence
460: relation. For details, including a very efficient algorithm to perform
461: minimization, see \cite{HU}.
462:
463:
464: \begin{definition}
465: A Moore machine $\A$ is called {\em counter-free\/} if there is no
466: word $w\in\Omega^+$ and no states $q_1,q_2,\dots,q_s,\ s>1,$ such that
467: $\hat\delta(q_1,w)=q_2,\dots,\hat\delta(q_{s-1},w)=
468: q_s,\hat\delta(q_s,w)=q_1.$
469: \end{definition}
470:
471:
472: Sometimes we use an extension of Moore machines---{\em Moore machines
473: with $\epsilon$-moves.} In a deterministic finite automaton an
474: $\epsilon$-move is a transition between two states done without
475: intervention of any letter from the input. By necessity, to maintain
476: the deterministic character of the automaton, an $\epsilon$-move must
477: not be combined with any other transitions starting from the same
478: state. For Moore machines, we adopt the convention that after
479: performing an $\epsilon$-move, no symbol is appended to the output.
480:
481: E.g., for the Moore machine with $\epsilon$-moves $\A$ below
482:
483:
484: \[\UseTips
485: \xymatrix
486: {
487: {}\ar[r]&
488: *+++[o][F]{\bot}
489: \ar@(dr,dl)[]^{\varrho}
490: \ar[r]^\omega&
491: *+++[o][F]{0}
492: \ar[r]^{\epsilon}&
493: *+++[o][F]{1}
494: \ar@(dr,ur)[]_{\varrho}
495: \ar@/_1cm/[ll]_{\omega}
496: }
497: \]
498:
499:
500: we have $f_{\A}(\omega\omega\omega)=0\bot0$ and
501: $f_{\A}(\omega\varrho\omega)=01\bot.$
502:
503: It is known that any function computable by a Moore machine with
504: $\epsilon$-moves can be computed by a Moore machine without
505: $\epsilon$-moves. Thus using $\epsilon$-moves we do not achieve
506: greater generality. However, some transformations of the machines can
507: be very conveniently represented by introducing $\epsilon$-moves.
508:
509:
510: \bigskip
511:
512: For us, Markov chains are a synonym of {\em Markov chains with
513: stationary transitions and finite state space.}
514:
515: Formally, given a finite set $I$ of {\em states} and a fixed function
516: $p:I\times I\to[0,1]$ satisfying \( (\forall i\in I)\qquad\sum_{j\in
517: I}p(i,j)=1, \) the {\em Markov chain\/} with state space $I$ and
518: transitions $p$ is a sequence $\X=X_0,X_1,\dots$ of random variables
519: $X_n:W\to I$, such that
520:
521: \begin{equation}\label{M1}\Pr(X_{n+1}=j|X_n=i)=p(i,j).\end{equation}
522:
523: The standard result of probability theory is that there exists a
524: probability triple $(W,\fM,\Pr)$ and a sequence $\X$ such that
525: \eqref{M1} is satisfied. $W$ is indeed the space of infinite sequences
526: of ordered pairs of elements from $I,$ and $\Pr$ is a certain product
527: measure on this set.
528:
529:
530: One can arrange the values $p(i,j)$ in a matrix $\Pi=(p(i,j);i,j\in
531: I).$ Of course, $p(i,j)\geq 0$ and $\sum_{j\in I}p(i,j)=1$ for every
532: $i.$ Every real square matrix $\Pi$ satisfying these conditions is
533: called {\em stochastic.}
534:
535: The initial distribution of $\X$ is that of $X_0,$ which can be
536: conveniently represented by a vector $\Xi_0=(p(i); i\in I).$
537: Its choice is independent from the function $p(i,j).$
538:
539:
540: It follows by a simple calculation that
541:
542: \begin{equation}\label{MP}
543: \Pr(X_{n+1}=j|X_n=i_n,X_{n-1}=i_{n-1},\dots,X_1=i_1)
544: =\Pr(X_{n+1}=j|X_n=i),
545: \end{equation}
546: which is called the {\em Markov property.}
547:
548: For our purposes, it is convenient to imagine the Markov chain $\X$ in
549: another, equivalent form: Let $K_I$ be the complete directed graph on
550: the vertex set $I.$ First we randomly choose the starting vertex in
551: $I,$ according to the initial distribution. Next, we start walking in
552: $K_I;$ at each step, if we are in the vertex $i,$ we choose the edge
553: $(i,j)$ to follow with probability $p(i,j).$ If we define
554: $X_n=(\text{the vertex in which we are after $n$ steps}),$ then $X_n$
555: is indeed the same $X_n$ as in \eqref{M1}.
556:
557: So we will be able to {\em draw\/} Markov chains. Doing so, we will
558: often omit edges $(i,j)$ with $p(i,j)=0.$
559:
560:
561: \subsection{Conditional objects and conditional events}
562:
563: Let for $(\varphi|\psi)\in\TT$ the function
564: $c=c_{(\varphi|\psi)}:\Omega^+\to\trzy$ be defined by
565:
566: \[c(w)=\begin{cases}1&\text{if $(\varphi|\psi)$ is true in $(w,n)$}\\
567: 0&\text{if $(\varphi|\psi)$ is false in $(w,n)$}\\
568: \bot&\text{if $(\varphi|\psi)$ is undefined in $(w,n)$}
569: \end{cases}
570: \]
571:
572: $\CC$ is the set of all functions $c:\Omega^+\to\trzy$ definable in
573: $\TT,$ and $\CC_+$ is the set of all functions
574: $c_+:\Omega^+\to\trzy^+$ computable by counter-free $\trzy$-valued
575: Moore machines.
576:
577: $\CC$ and $\CC_+$ are isomorphic under the mapping $\CC_+\ni c_+\mapsto
578: c\in\CC$ defined by
579:
580: \[c(w)=\text{last-letter-of}(c_+(w)).\]
581:
582: The sets $\CC$ and $\CC_+$ are regarded as two representations of {\em
583: conditional objects}. We have yet another representation, denoted
584: $\CC_\infty$: it consists of functions $\Omega^\infty\to\trzy^\infty,$
585: and $c_\infty,$ the third representation of the same conditional, is an
586: infinite sequence of values of $c$ on all finite nonempty prefixes of
587: $w.$
588:
589: We will be using the name {\em conditional events\/} to refer to
590: conditionals considered with a probability space in the background.
591: \bigskip
592:
593:
594: \begin{definition}[Conditional event \cite{TT}] Let $c\in \CC$ be a
595: conditional object over $\Omega$ and let $\A$ be any counter-free
596: Moore machine with output function $h,$ computing $c_+.$ Suppose
597: $\Omega$ is endowed with a probability space structure
598: $(\Omega,\fM,\Pr).$ The conditional object $c$ becomes then a sequence
599: $\ce{c}=\ce{c}_1,\ce{c}_2,\dots$ of random variables
600: $\ce{c}_n:\Omega^\infty\to\trzy,$ defined by the formula
601:
602: \begin{equation}\label{e1}
603: \ce{c}_n(w)=c(\text{prefix-of-length-$n$-of}(w)),
604: \end{equation}
605:
606: where $\Omega^\infty$ is considered with the product probability structure.
607:
608: We call $\ce{c}$ the {\em conditional event\/} associated with $c.$
609:
610: Moreover, in presence of probability space structure $\A$ becomes a
611: Markov chain $\X(\A)$ (by replacing labels of th e transitions by
612: their probabilities under $\Pr$ in the diagram of $\A$), and then
613: $\ce{c}=h(\X),$ where $h$ is the output function of $\A.$
614: \end{definition}
615:
616: In particular, $\Pr(\ce{c}_n=\true)$ is the probability that at time
617: $n$ the conditional object $c$ is true, $\Pr(\ce{c}_n=\false)$ is the
618: probability that at time $n$ the conditional object $c$ is false, and
619: $\Pr(\ce{c}=\bot)$ is the probability that at time $n$ the conditional
620: object $c$ is undefined.
621:
622:
623: \begin{definition}[asymptotic probability \cite{TT}]
624: We define the {\em asymptotic probability at time $n$} of a
625: conditional event $c\in\CC$ by the formula
626:
627: \begin{equation}\label{e2}
628: \pr_n(c)=\dfrac{\Pr(\ce{c}_n=1)}
629: {\Pr(\ce{c}_n=0\ \text{or}\ 1)}.
630: \end{equation}
631:
632: If the denominator is $0,$ $\pr_n(c)$ is undefined.
633:
634: The {\em asymptotic probability\/} of $c$ is
635:
636: \begin{equation}\label{e3.5}
637: \Pr(c)=\lim_{n\to\infty}\pr_n(c),
638: \end{equation}
639:
640: provided that $\pr_n(c)$ is defined for all sufficiently large $n$ and
641: the limit exists.
642:
643: If $\varphi\in\TL$ then we write $\Pr(\varphi)$ for $\Pr((\varphi|\TRUE)).$
644: \end{definition}
645:
646: \begin{theorem}[Bayes' Formula \cite{TT}]\label{Bayes}
647:
648: Let $(\varphi|\psi)$ be a conditional object over $\Omega$ endowed
649: with a probability space structure $(\Omega,\fM,\Pr),$ and let $\A$ be
650: any counter-free Moore machine with output function $h,$ computing
651: $(\varphi|\psi).$
652: \begin{enumerate}
653: \item For every state $i$ of $\X(\A)$ the probability
654: $\lim_{n\to\infty}\Pr(X_n=i)$ exists.
655: \item For every $\star\in\trzy$ the probability
656: $\lim_{n\to\infty}\Pr(\ce{(\varphi|\psi)}_n=\star)$ exists.
657: \item The Bayes' Formula
658: \end{enumerate}
659:
660: \[\Pr((\varphi|\psi))=\frac{\Pr(\varphi\land\psi)}{\Pr(\psi)}\]
661:
662: holds whenever the right-hand-side above is well-defined, i.e.,
663: $\Pr(\psi)>0.$
664: \end{theorem}
665:
666: \section{Connectives of conditionals}
667:
668: If we wish to extend the classical two-valued conjunction to
669: conditionals, we are faced with the problem of {\em
670: synchronization}. Indeed, what is easy in $\dwa$-valued world becomes
671: messy in $\trzy $-valued world. The problem is that the conditionals
672: need not become defined synchronously. For the classical conjunction
673: this problem does not exist, because both arguments are always
674: defined. Now we have to resolve the question how to define the
675: conjunction, when some of the arguments are undefined.
676:
677:
678: \subsection{Present tense connectives}\label{teraz}
679:
680: Let us recall that present tense connectives are those, whose definition
681: in $\TT$ does not use temporal connectives, and therefore depends
682: on the present, only. Equivalently, an $n$-ary present tense connective is
683: completely characterized by a function $\trzy^n\to\trzy .$
684:
685: Here are several possible choices for the conjunction, which is always
686: defined as a pointwise application of the following $\trzy$ valued
687: functions. Above we display the notation for the corresponding kind of
688: conjunction.
689:
690: \[\label{toto}
691: \begin{array}{ccc}\conn{\land_{\SAC}}00001101&
692: \conn{\land_{\GNW}}00001\bot0\bot&
693: \conn{\land_{\Sch}}00\bot01\bot\bot\bot\\
694: &&\\
695: &{\begin{array}{|c|c|}\hline
696: \multicolumn{2}{|c|}{\lnot x} \\\hline\hline x&\lnot x\\\hline 0&1\\
697: 1&0\\\bot&\bot\\\hline\end{array}}&\\ &&\\\conn{\lor_{\SAC}}01011101&
698: \conn{\lor_{\GNW}}01\bot111\bot1&
699: \conn{\lor_{\Sch}}01\bot11\bot\bot\bot.
700: \end{array}\]
701:
702:
703:
704: They can be equivalently described by syntactical manipulations in $\TT.$
705: The reduction rules are as follows:
706:
707:
708: \begin{equation}\label{TLTL}
709: \begin{split}
710: (a|b)\land_\SAC(c|d)&=(abcd\lor abd^\C\lor cdb^\C|b\lor d)\\
711: (a|b)\land_\GNW(c|d)&=(abcd|a^\C d\lor c^\C d\lor abcd)\\
712: (a|b)\land_\Sch(c|d)&=(abcd|bd)\\
713: \llnot_0 (a|b)&=(a^\C|b)\\
714: (a|b)\lor_\SAC(c|d)&=(ab\lor cd|b\lor d)\\
715: (a|b)\lor_\GNW(c|d)&=(ab\lor cd|ab\lor cd \lor bd)\\
716: (a|b)\lor_\Sch(c|d)&=(ab\lor cd|bd).
717: \end{split}
718: \end{equation}
719:
720: The first is based on the principle ``if any of the arguments becomes
721: defined, act!''. A good example would be a quotation from \cite{c97}:
722:
723: \begin{quote}\sl ``One of the most dramatic examples of the unrecognized
724: use of compound conditioning was the first military strategy of our
725: nation. As the Colonialists waited for the British to attack, the
726: signal was `One if by land and two if by sea'. This is the conjunction
727: of two conditionals with uncertainty!''\footnote{NB, if the British had
728: decided to attack from both directions, but not simultaneously, we would
729: have probably discovered temporal conditionals much earlier.}
730: \end{quote}
731:
732: Of course, if the above was understood as a conjunction of two
733: conditionals, the situation was crying for the use of $\land_\SAC,$
734: whose definition has been proposed independently by Schay, Adams and
735: Calabrese (the author of the quotation).
736:
737: The conjunction $\land_{\GNW}$ represents a moderate approach, which
738: in case of an apparent evidence for $\false$ reports $\false,$ but
739: otherwise it prefers to report unknown in a case of any doubt. Note
740: that this conjunction is essentially the same as {\em lazy
741: evaluation}, known from programming languages.
742:
743: Finally, the conjunction $\land_\Sch$ is least defined, and acts
744: (classically) only if both arguments become defined. It corresponds to
745: the {\em strict evaluation}.
746:
747: We have given an example for the use of $\land_\SAC.$ The uses of
748: $\land_\GNW$ and $\land_\Sch$ can be found in any computer program
749: executed in parallel, which uses either lazy or strict evaluation of
750: its logical conditions. And indeed both of them happily coexist in
751: many programming languages, in that one of them is the standard
752: choice, the programmer can however explicitly override the default and
753: choose the other evaluation strategy.
754:
755: This seems to suggest that neither of the three choices discussed in
756: this paragraph is {\em the\/} conjunction of conditionals. There are
757: indeed many possible choices, and all of them have their own merits.
758: And indeed already the original system of Schay consisted of five
759: operations: $\llnot_0,\land_\SAC,\lor_\SAC,\land_\Sch$ and $\lor_\Sch.$
760: Moreover, he was aware that these operations still do not make the
761: algebra functionally complete (even in the narrowed sense, restricted to
762: defining only operations which are undefined for all undefined
763: arguments). And in order to remedy this he suggested to use one of
764: several additional operators, one of them being $\land_\GNW!$ So for him
765: all those operations could coexist in one system.
766:
767: \paragraph{Present tense re-conditioning}
768:
769: Calabrese \cite{C90} and Goodman, Nguyen and Walker \cite{gn96} proposed
770: their own extensions of the conditioning operator to $\trzy,$ hence
771: making it available for re-conditioning in $\SAC$ and $\GNW$,
772: respectively. The definitions are
773:
774: \[ \conn{|_\SAC}\bot\bot\bot01\bot01\ \ \
775: \conn{|_\GNW}\bot\bot\bot01\bot0\bot\]
776:
777:
778:
779:
780: \subsection{Past tense connectives}
781:
782: Now we consider connectives, whose definitions refer to the strict
783: past of their arguments. We continue to consider conjunction, which we
784: use as a kind of model example.
785:
786:
787: \paragraph{Examples of past tense connectives}
788:
789: The following are connectives very close to the conjunction and
790: disjunction of the {\em product space\/} \cea\/ introduced in
791: \cite{g94}\footnote{See the discussion on embedding \cea's in our model
792: below.}, defined by the rule: the conjunction is defined and true iff both of
793: its arguments have been defined, and moreover the historically {\em first\/}
794: values of its two arguments have been both $1.$ Otherwise it is defined and
795: false. Disjunction is defined similarly. They use the ``Russian roulette''
796: approach to repeating experiments.
797:
798: In the language of $\TT$ $(a|b)\land_\PS (c|d)$ can be expressed by
799:
800:
801: \begin{equation}\label{land_PS}
802: \left(\left.
803: \begin{array}{c}
804: \PDIA(a\land b\land
805: (\lnot\PREV\PDIA b))\\
806: \land\\
807: \PDIA(c\land d\land
808: (\lnot\PREV\PDIA d))
809: \end{array}
810: \right|
811: \TRUE
812: \right),
813: \end{equation}
814:
815: and the definition of and $(a|b)\lor_\PS(c|d)$ is similar. They seem
816: complicated, but can be simplified, what we do below, and the Moore
817: machine representations are again much simpler and easier to analyze.
818:
819: \begin{figure}[!hp]
820:
821: \[\UseTips
822: \xymatrix @C=20mm @R=12mm
823: {
824: &&*+++[o][F]{0}
825: \ar@(ul,ur)[]
826: \ar[d]^{\textstyle{dc^\C}}
827: \ar[rd]^{\textstyle{dc}}\\
828: {}\ar[r]&*+++[o][F]{0}
829: \ar@(u,l)[]
830: \ar[ur]^{\textstyle{ba}}
831: \ar[dr]_{\textstyle{dc}}
832: \ar@<1ex>[r]^(.65){\textstyle{ba^\C}}
833: \ar@<-1ex>[r]_(.65){\textstyle{dc^\C}}
834: \ar@/_3cm/[rr]_{\textstyle{abcd}}
835: &*+++[o][F]{0}
836: \ar@(ur,dr)[]
837: &*+++[o][F]{1}
838: \ar@(ur,dr)\\
839: &&*+++[o][F]{0}
840: \ar@(dl,dr)[]
841: \ar[u]_{\textstyle{ba^\C}}
842: \ar[ru]_{\textstyle{ba}}
843: }
844: \]
845:
846: \caption[PS conjunction]{Moore machine of $\ps{(a|b)\land(c|d)}$}\label{f3}
847: \end{figure}
848:
849:
850: To simplify the $\TT$ representation above, let us define
851: $\first(a|b)$ to be
852:
853: \begin{equation}\label{first}\first(a|b):=(\PDIA(a\land b\land
854: \lnot\PREV\PDIA b))|\TRUE ).
855: \end{equation}
856:
857: It is convenient to denote by $\first_\dwa(a|b)$ the first argument of
858: $\first(a|b).$
859:
860: Then the conjunction can be easily and effectively described as
861: $(\first_\dwa(a|b)\land\first_\dwa(c|d)|\TRUE)$ (where $\land$ is the
862: classical conjunction of temporal logic).
863:
864: The minimal Moore machine of $\first(a|b)$ is depicted in Fig.\
865: \ref{f4} below.
866:
867:
868: \begin{figure}[!hp]
869:
870: \[\UseTips
871: \xymatrix @R=4mm
872: {
873: &&*+++[o][F]{1}
874: \ar@(ur,dr)[]\\
875: {}\ar[r]&*+++[o][F]{0}
876: \ar@(ur,dr)[]^{\textstyle{b^\C}}
877: \ar@/^/[ur]^{\textstyle{ba}}
878: \ar@/_/[dr]_{\textstyle{ba^\C}}\\
879: &&*+++[o][F]{0}
880: \ar@(ur,dr)[]
881: }
882: \]
883:
884: \caption[PS simple conditional]{Moore machine of $\ps{(a|b)}=\first(a|b).$}\label{f4}
885: \end{figure}
886:
887:
888:
889:
890:
891: \section{Embedding of existing {\em cea\/}'s and their incompleteness}
892:
893:
894: In this section we want to discuss the problem of embedding existing
895: \cea's into our model, and on that basis, the problem of defining
896: natural connectives among conditionals in general.
897:
898: \subsection{Syntax}
899:
900: We assume the following syntax of the {\em flat conditional
901: expressions}. The set of all such expressions will be denoted $\LL.$
902:
903: The set of these expressions is the smallest set, containing all {\em
904: simple conditionals\/} of the form $(x|y),$ where $x,y\in\Omega,$ and
905: closed under two-ary (infix) operations $\land,\lor$ and one unary
906: prefix operation $\llnot.$
907:
908: If we require the closure under one additional binary operation
909: $(\cdot|\cdot)$ (which shouldn't be mixed up with the
910: parenthesis-bar-parenthesis construction appearing in simple
911: conditionals), we obtain the set of {\em full conditional expressions},
912: denoted $\LL^|.$
913:
914: \subsection{Conditional event algebras}\label{cea}
915:
916:
917:
918: According to \cite{kniga}, a {\em conditional event algebra\/} (\cea{}
919: in short) over a probability space $(\Omega,\fM,\Pr)$ is a space (but
920: not necessarily a probability space) $(\Omega_o,\fM_o,\Pr_o),$ extending
921: $(\Omega,\fM,\Pr),$ together with a function
922: $(\cdot|\cdot):\fM\times\fM\to\fM_o$ such that
923: \begin{itemize}
924: \item $\fM_o$ is an algebra of the signature of boolean algebras.
925: \item $(a|b)=(a\cap b|b)$ for all $a,b\in \fM.$
926: \item The subalgebra of $(\Omega_o,\fM_o,\Pr_o)$ consisting of the
927: elements $(a|\Omega)$ is isomorphic to $(\Omega,\fM,\Pr)$ under the
928: bijection $a\mapsto(a|\Omega).$
929: \item $\Pr_o((a|b))=\Pr(a\cap b)/\Pr(b)$ for $a,b\in\fM,\ \Pr(b)>0.$
930: \item Certain equalities hold among the $\Pr_o$-probabilities.
931: \item The $\Pr_o$-probabilities for $\cap,\cup$ and ${}^\C$ of elements
932: of the form $(a_i|b_i)$ for $a_i,b_i\in\fM$ are effectively computable
933: from the set of $\Pr$-probabilities of all the boolean combinations of
934: the elements $a_i,b_i.$
935: \end{itemize}
936:
937: \paragraph{How we understand \cea's.} It is readily seen, that any
938: particular \cea{} over any $(\Omega,\fM,\Pr)$ can be equivalently
939: considered as a mapping assigning elements of $\fM_o$ and probabilities
940: to flat conditional expressions. In this sense, \cea{} is a kind of a
941: {\em model\/} of a logic, whose syntax are the flat conditional
942: expressions. This is the way we understand \cea's and this is the level
943: on which we will criticize them.
944:
945: \paragraph{Probabilistic models and \cea's.}
946:
947: First of all, we do not think that the algebraic structure of a \cea{}
948: is particularly important. The language of flat conditional
949: expressions does not have equality, so what is really crucial are the
950: probabilities. The algebraic structure can indeed be an obstacle
951: while assigning probabilities (this is perhaps why the authors of the
952: earlier papers devoted so much attention to it), but otherwise we are
953: not so much interested in it.
954:
955: Next, what a \cea{} assigns to a conditional expression is definitely too
956: little. Apart from an element of $\fM_0$ and probability, an experiment
957: should be determined to verify experimentally the value of the
958: probability. This means, that the objects assigned to conditional
959: expressions should be events in a probabilistic space, i.e., the triple
960: $(\Omega_o,\fM_o,\Pr_o)$ should be indeed a probability space.
961: Moreover, the experiments for simple conditionals $(a|b)$ should
962: correspond to the natural experiments one performs to learn conditional
963: probabilities. In such experiments one can typically measure another
964: probabilistic parameters, like, e.g., the probability that $(a|b)$ is
965: defined. We think such additional parameters should be assigned to
966: compound conditional events, too. Of course, in the existing algebras
967: we have hints concerning it, hidden in their universes and other details
968: of the constructions, provided by the inventors, but the very definition
969: of a \cea{} does not require the additional parameters to be even defined,
970: let alone to satisfy any reasonable properties. Strictly speaking, the
971: signature of \cea's is too small. In its present shape it permits
972: existence of other, isomorphic algebras, where all the additional
973: information is lost.
974:
975:
976: Last, but not least, in the applications of classical probability theory
977: one often encounters problems in modeling, typically of the following
978: form: one has an event, whose meaning is completely clear (it is known,
979: when it happens and when it doesn't), but there is a problem of
980: specifying the probability space structure, and sometimes different
981: choices lead to different values for the probability of the same event.
982: In such circumstances one can only experimentally decide which of the
983: models is the correct one. A standard example is the difference between
984: the so called statistics of: Maxwell-Boltzmann, Bose-Einstein and
985: Fermi-Dirac, considered in quantum physics, and the unsuccessful search
986: for any elementary particle, which would satisfy the first statistics,
987: seemingly the most natural one among them (therefore we have bosons and
988: fermions, but we do not have maxwellons in physics) \cite{feller1}. The
989: tremendous success of probability theory in applications seems to
990: suggest this is the right way of creating a mathematical model of a real
991: life situation. However, the \cea{} offers us another challenge. Even
992: after coming up with the proper model of probability space of
993: unconditional events, one still has a lot of work to choose the right
994: model for conditional events.
995:
996: In plain words, it means that the structure modeling conditional events
997: over a given probabilistic space of unconditional events, should be
998: functorial: given the former space, the space of conditional events
999: ought to be uniquely determined. Here the \cea's again fall short of
1000: satisfying this requirement, because there are many known \cea's, and
1001: each of them has its own definition of $(a|b)\land(c|d),$ with its own
1002: probability, and all of them derive their definitions from certain first
1003: principles. The answer is almost obvious --- the signature is too small,
1004: and should consist of many different conjunctions, disjunctions and
1005: negations, and perhaps lots of other connectives, which do not have any
1006: natural counterparts in the nonconditional world.
1007:
1008: \paragraph{$\TT$ as a solution.}
1009:
1010: We believe that our system of temporal conditional events addresses all
1011: of the problems we have indicated above.
1012:
1013: Our conditional events belong to a normal probability space. They are
1014: indeed stochastic processes --- projections of Markov chains, and all
1015: their probabilistic properties (many more than just the bare
1016: probability) can be verified experimentally,
1017:
1018: To the contrary, all the \cea's we call present tense in this paper have
1019: a natural representation as algebras of $\trzy $-valued indicator
1020: functions \cite[Chapter 3]{gnw91}. Recalling that every event from
1021: $\Sigma$ can be equivalently characterized by its $\dwa$-valued {\em
1022: indicator random variable}, which is nothing but the characteristic
1023: function, we should naturally expect that the lifting of $\Pr$ to the
1024: space of $\trzy$-valued indicator functions, should consist of {\em
1025: $\trzy $-valued random variables}, while the definition of a \cea{}
1026: requires just elements of a strange algebra $\fM_o$ and numbers. In
1027: case of our temporal conditionals, we naturally expect conditionals to
1028: be functions $\Omega^\infty\to \trzy ^\infty,$ and hence the lifting of
1029: $\Pr$ to consist of random functions, which are nothing but {\em
1030: stochastic processes}, the choice we actually have made. This is the
1031: natural pattern one should follow, and this is where the scalability of
1032: the model is hidden. E.g., nothing could prevent us from considering
1033: continuous time stochastic processes as models of conditional events, if
1034: need be.
1035:
1036:
1037: The space of conditional events is uniquely determined by the
1038: probabilistic space of nonconditional events, and has many more natural
1039: connectives than just three. In fact, we are not very original here:
1040: already Schay \cite{MR37:5901} proposed a system with five connectives,
1041: and considered adding even more of them.
1042:
1043: Finally, to make our argument complete, we consider the major \cea's
1044: below, and show that they embed in our models, in many different
1045: ways. For the product space \cea, we employ the embeddings to
1046: demonstrate that this \cea{} has some further defects. The existence
1047: of multiple embeddings shows that the structure of a \cea's isn't
1048: functorial just because there are many known \cea's, but because the
1049: structure of a \cea{} does not prescribe the way the conditionals
1050: relate to nonconditional events. Additionally, the product space
1051: \cea{} is unable to characterize independence of conditional events by
1052: means of equalities of probabilities.
1053:
1054: \subsection{Present tense \cea's}
1055:
1056:
1057: In our framework, there is a distinctive class of \cea's, which we call
1058: {\em present tense \cea's.} The definitions of their connectives refer
1059: to the present of the process, only, hence the name. Consequently, such
1060: a connective applied to simple conditionals yields simple conditionals,
1061: again. Among such systems are SAC, proposed independently by Schay
1062: \cite{MR37:5901}, Adams \cite{a86} and Calabrese \cite{c87} (who
1063: extended it by an operator for re-conditioning), and GNW proposed by
1064: Goodman \cite{g87} and Goodman, Nguyen and Walker \cite{gnw91} (later
1065: Goodman and Nguyen \cite{gn96} proposed a re-conditioning operator for
1066: GNW). They assume that all boolean combinations of simple conditional
1067: expressions yield simple conditional expressions again. The definitions
1068: of their connectives are given in Section \ref{teraz}. They do form
1069: \cea's. Equivalently, \cite[Chapter 3]{gnw91}, these algebras can be
1070: characterized as algebras of $\trzy$-valued indicator functions, and
1071: their connectives are then characterized by mappings from certain
1072: Cartesian power of $\trzy$ into $\trzy.$
1073:
1074: We take the second point of view, and define their semantics as
1075: follows. Our definition assigns to every conditional expression $e$ a
1076: $\trzy$-valued indicator function $\sac{e}$ and $\gnw{e},$ respectively.
1077: For us, indicator functions are nothing else than present tense
1078: conditionals. (Originally these algebras do not involve time.) So, we
1079: give the definition by describing translations
1080: $\sac{\cdot},\gnw{\cdot}:\LL\to\CC.$ We use in the translations present
1081: tense connectives of conditionals defined in \eqref{TLTL}.
1082:
1083: The definition of SAC:
1084:
1085: \begin{equation}\label{sac}
1086: \begin{split}
1087: \sac{(a|b)}&=(a|b),\\
1088: \sac{e\land e'}&=\sac{e}\land_\SAC\sac{ e'},\\
1089: \sac{e\lor e'}&=\sac{e}\lor_\SAC\sac{ e'},\\
1090: \sac{\llnot e}&=\llnot_0\sac{e},\\
1091: \sac{(e| e')}&=(\sac{e}|_\SAC\sac{ e'}).
1092: \end{split}
1093: \end{equation}
1094:
1095: The definition of GNW:
1096: \begin{equation}\label{gnw}
1097: \begin{split}
1098: \gnw{(a|b)}&=(a|b),\\
1099: \gnw{e\land e'}&=\gnw{e}\land_\GNW\gnw{ e'},\\
1100: \gnw{e\lor e'}&=\gnw{e}\lor_\GNW\gnw{ e'},\\
1101: \gnw{\llnot e}&=\llnot_0\gnw{e},\\
1102: \gnw{(e| e')}&=(\gnw{e}|_\GNW\gnw{ e'}).
1103: \end{split}
1104: \end{equation}
1105:
1106: Given a probability space $(\Omega,\PP(\Omega),\Pr),$ the probability of
1107: a conditional expression $e$ is
1108: $\Pr_\SAC(e)={\Pr(\sac{e}=1)}/{\Pr(\sac{e}\neq \bot)},$ and
1109: similarly $\Pr_\GNW(e)={\Pr(\gnw{e}=1)}/{\Pr(\gnw{e}\neq \bot)},$
1110: provided that the denominators are nonzero.
1111:
1112: All theses systems are readily seen to embed in our system of
1113: conditionals. In fact, if one represents them in the form of reduction
1114: rules, as in \eqref{TLTL}, they do even embed syntactically in the $\TT$
1115: logic. They are present tense because they do not contain temporal
1116: connectives.
1117:
1118: \subsection{Product space \cea}
1119:
1120: However, there is another \cea, called the {\em product space} \cea,
1121: which is not present tense. In order to analyze it and show that it can
1122: be interpreted in our model, we have to give the definition.
1123:
1124:
1125: The semantics is as follows:
1126:
1127: Beginning with $(\Omega,\fM,\Pr),$ we form its countable power
1128: $\Omega^\infty$ endowed with the product measure. The cylinder
1129: $\underbrace{b\times\dots\times b}_j\times
1130: a\times\Omega\times\Omega\times\cdots\subseteq \Omega^\infty$ for
1131: $a,b\in\Sigma$ is denoted $b^j\times a\times\hat{\Omega}.$
1132:
1133: Define the semantics function $\ps{\cdot}:\PS\to\PP(\Omega^\infty)$ by
1134:
1135:
1136: \begin{align*}
1137: \ps{(a|_\PS b)}&=
1138: \bigcup_{i=0}^\infty (\Omega\setminus b)^j\times (b\cap
1139: a)\times\hat{\Omega},\\
1140: \ps{ e\land e'}&=\ps{ e}\cap\ps{ e'},\\
1141: \ps{ e\lor e'}&=\ps{ e}\cup\ps{ e'},\\
1142: \ps{\llnot e}&=\Omega^\infty\setminus\ps{ e}.
1143: \end{align*}
1144:
1145: $\fM_o$ of the product space \cea\ is then the subalgebra of the
1146: (boolean) algebra $\langle\PP(\Omega^\infty),\cup,\cap,
1147: (\Omega^\infty\setminus \cdot)\rangle,$ generated by all elements
1148: $b^j\times a\times\hat{\Omega}$ where $a,b\in\Sigma,$ and $\Pr_o$ is the
1149: product measure.
1150:
1151: There are indeed two versions of $\PS$: one defined in the paper
1152: \cite{g94}, where equality of two conditionals is understood as true
1153: equality of sets, and another, defined in \cite{kniga}, where the
1154: equality of conditionals is understood as {\em equality almost
1155: everywhere}, i.e., two conditional events of $\PS$ are equal iff their
1156: symmetric difference has probability $0$. The latter is therefore not
1157: logical, since it depends on the particular probability space structure.
1158:
1159: The probabilities assigned to the elements of $\PS$ are those according
1160: to the infinite product of $\Pr.$
1161:
1162: \subsection{First embedding}\label{FE}
1163:
1164: In order to construct the first embedding of $\PS$ into $\CC$ by
1165: defining two operations $\ttl{\cdot}:\LL\to\TL$ and
1166: $\ttt{\cdot}:\LL\to\TT$ as follows:
1167:
1168: \begin{equation}\label{pierwszy}
1169: \begin{split}
1170: \ttl{(a| b)}&=\first_\dwa(a|b)\\
1171: \ttl{e \land e'}&=\ttl{e}\land\ttl{e'}\\
1172: \ttl{e \lor e'}&=\ttl{e}\lor \ttl{e'}\\
1173: \ttl{\llnot e}&=\lnot\ttl{e}\\
1174: \ttt{e}&=(\ttl{e}|\TRUE).
1175: \end{split}
1176: \end{equation}
1177:
1178: $\ttt{e}$ (or, more formally, the conditional from $\CC$ represented by
1179: the former) is the desired embedding.
1180:
1181: \begin{lemma}\label{cztery-lematy}
1182: For every expression $e\in\LL$
1183:
1184: \begin{enumerate}
1185: \item For every word $w\in\Omega^\infty$ holds
1186: $(\ttt{e})_\infty(w)\in\dwa^\infty;$
1187:
1188: \item Suppose $(a_1|b_1),\dots,(a_n|b_n)$ are all simple conditionals
1189: occurring in $e.$ Suppose $w\in\Omega^\infty$ is so that
1190: $b_{i_1},\dots,b_{i_k}$ are all events among $b_1,\dots,b_n$ which
1191: happen in the sequence $w,$ and all of them happen not later than at
1192: time $m.$ Then of the word $(\ttt{e})_\infty(w)$ is constant beginning
1193: since time $m.$
1194:
1195: \item $\ps{e}=\{w\in\Omega^\infty~/~\text{\rm $(\ttt{e})_\infty(w)$ is
1196: eventually constant 1}\}$;
1197:
1198: \item $\Pr_o(e)=\Pr(\ttt{e})$;
1199: \end{enumerate}
1200: \end{lemma}
1201: \begin{proof} The proof of 1., 2.\ and 3.\ goes by simultaneous
1202: induction w.r.t.\ $e.$ For $e=(a| b)$ they follows from
1203: a simple analysis of the definition of $\ps{e}$ and $\first_\dwa(a|b).$
1204:
1205:
1206: Now consider $e=\lnot e'$ and assume by induction that 1., 2.\ and 3.\
1207: hold for $e'.$ Since $\ttl{\lnot e}=\lnot\ttl{e'},$ we have 1.\ and 2.\
1208: immediately.
1209:
1210: Moreover, $(\lnot\ttl{e'}|\TRUE)_\infty(w)$ is eventually constant 1 iff
1211: $(\ttl{e'}|\TRUE)_\infty(w)$ is eventually constant $0,$ which by 1.\
1212: and 2.\ for $e'$ is equivalent to the fact that
1213: $(\ttl{e'}|\TRUE)_\infty(w)$ is not eventually constant 1. This
1214: concludes the induction step of 3.
1215:
1216: Induction steps for the other connectives are equally simple.
1217:
1218: We turn now to 4. Suppose $(a_1|b_1),\dots,(a_n|b_n)$ are all simple
1219: conditionals occurring in a conditional expression $\varphi.$ W.l.o.g.\
1220: assume $\Pr(b_i)>0$ for $i=1,\dots, k$ and $\Pr(b_i)=0$ for
1221: $i=k+1,\dots, n.$ (We permit $k=0$ and $k=n,$ in which cases either all
1222: $b_i$ are impossible, or all of them have positive probability.)
1223:
1224: Represent $\Omega^\infty$ as a disjoint union of sets
1225:
1226: \[A_n:=\{w\in\Omega^\infty~/~\begin{array}{c}
1227: \text{$b_1,\dots,b_k$ happen in
1228: $w$ and the first time}\\\text{when they all have already happened is
1229: $n$}\end{array}\}\]
1230:
1231: and the set $A_\infty:=\{w\in\Omega^\infty~/~\text{not all of
1232: $b_1,\dots,b_k$ happen in $w$}\}.$ All these sets are clearly
1233: measurable.
1234:
1235: It is not hard to see that the (product) probability of
1236: $A_\infty$ is $0,$ so
1237: $\Pr_o(e)=\sum_{\substack{n\in\N\\A_n\subseteq\ps{e}}}
1238: \Pr(A_n).$
1239:
1240: It is not hard to verify, either, that
1241:
1242: \[\lim_{n\to\infty}\sum_{m\in\N} \Pr\nolimits_n(A_m) \ =\ 1.\]
1243:
1244: These two equalities imply 4.\ immediately, since for every $n$
1245:
1246: \[\sum_{\substack{m\in\N\\A_m\subseteq\ps{e}}}
1247: \Pr\nolimits_n(A_m)\leq \Pr\nolimits_n(\ttt{e})\leq 1-
1248: \sum_{\substack{m\in\N\\A_m\cap\ps{e}=\varnothing}}
1249: \Pr\nolimits_n(A_m).\]
1250: \end{proof}
1251:
1252: The following theorem follows now instantly.
1253:
1254: \begin{theorem}\label{zaba} $\ttt{\cdot}$ is an embedding of the $\PS$
1255: \cea{} into $\CC,$ in the sense that for any underlying probability
1256: space, and any conditional expressions $e,e'$, $\ttt{e}=\ttt{e'}$
1257: iff $\ps{e}=\ps{e'},$ and $\Pr(\ttt{e})=\Pr_o(e).$\qed
1258: \end{theorem}
1259:
1260:
1261: \subsection{Reverse embeddings}
1262:
1263: \begin{definition}
1264: The reverse of a word $w=\omega_1\dots\omega_n\in \Omega^+,$ denoted
1265: $w^{\RR},$ is $\omega_n\omega_{n-1}\dots \omega_1.$
1266:
1267: Now consider a conditional $c\in\CC.$ Then $c^{\RR}\in\CC$ is a conditional
1268: defined by
1269:
1270: \[c^{\RR}(w):=c(w^{\RR}).\]
1271: \end{definition}
1272:
1273: The class of languages definable in $\TL$ is reverse-closed
1274: \cite{Emerson}, i.e., if $L=\{w~/~w,|w|\models\varphi\}$ for some
1275: $\varphi\in\TL$, then $L^{\RR}=\{w^{\RR}~/~w\in
1276: L\}=\{w~/~w,|w|\models\psi\}$ for some$\varphi\in\TL$. It follows
1277: that $c^{\RR}$ is indeed a conditional in our sense.
1278:
1279:
1280:
1281: \begin{theorem}\label{reverse}
1282: For every $\star\in\trzy$
1283: \[
1284: \Pr(\ce{c}_n=\star)=\Pr(\ce{c^{\RR}}_n=\star).
1285: \]
1286:
1287: Consequently, $\Pr_o(c)=\Pr(c^\RR).$
1288: \end{theorem}
1289: \begin{proof}
1290: $\Pr$ is understood here as a product measure, which is insensitive to
1291: the order of its coordinates.
1292: \end{proof}
1293:
1294:
1295:
1296:
1297: It follows that any \cea, which can be at all isomorphically embedded
1298: in our stochastic process model, has at least {\em two\/} embeddings,
1299: which are reverses of each other. The only exception is when the
1300: embedding is invariant under reverse, i.e., when each conditional $c$
1301: in the image of the embedding satisfies $c(w)=c(w^\RR)$ for all
1302: $w\in\Omega^+.$ However, it seems unlikely that any reasonable
1303: embedding has this property. In particular, the natural embeddings of
1304: the \cea's we consider here are not of this kind.
1305:
1306: As a matter of example, we consider here $\PS$. For the $\PS$
1307: conjunction, its informal description of its reverse representation in
1308: $\CC$ is that it is always defined and true iff the most recent defined
1309: values of its both arguments were $1.$
1310:
1311:
1312: In $\TT,$ we have that $\ttr{(a|b)\land(c|d)}$ (the reverse of the
1313: embedding $\ttt{\cdot}$ defined in \eqref{pierwszy}) is defined by
1314: $((b^\C\Since (a\land b))\land(d^\C\Since (c\land d))|\TRUE),$ whose
1315: Moore machine is depicted below.
1316:
1317:
1318: \begin{figure}[!hbp]
1319:
1320: \[\UseTips
1321: \xymatrix @C=16mm @R=36mm
1322: {
1323: &&*+++[o][F-]{0}
1324: \ar@(ul,ur)
1325: \ar@<2ex>[dll]|(.36){c^\C d\land (ab)^\C}
1326: \ar@<2ex>[ddl]|(.36){abc^\C d}
1327: \ar@<2ex>[dr]|(.36){ab\land (c^\C d)^\C}
1328: \\
1329: *+++[o][F-]{0}
1330: \ar@(l,u)
1331: \ar@<2ex>[rrr]|(.36){abcd}
1332: \ar@<2ex>[rru]|(.36){cd\land (ab)^\C}
1333: \ar@<2ex>[dr]|(.36){ab\land (cd)^\C}
1334: &&&*+++[o][F-]{1}
1335: \ar@(ur,dr)
1336: \ar@<2ex>[lll]|(.36){a^\C bc^\C d}
1337: \ar@<2ex>[lld]|(.36){c^\C d\land (a^\C b)^\C}
1338: \ar@<2ex>[lu]|(.36){a^\C b\land (c^\C d)^\C}
1339: \\
1340: {}\save[]+<-2cm,3cm>\ar[u]\restore %strzalka rozpoczynajaca rysunek
1341: &*+++[o][F-]{0}
1342: \ar@(rd,dl)
1343: \ar@<2ex>[rru]|(.36){cd\land(a^\C b)^\C}
1344: \ar@<2ex>[uur]|(.36){a^\C bcd}
1345: \ar@<2ex>[ul]|(.36){a^\C b\land (ab)^\C}
1346: }
1347: \]
1348:
1349: \caption[Reverse conjunction]{Moore machine of $\ttr{(a|b)\land(c|d)}.$}
1350: \end{figure}
1351:
1352:
1353: The precise definition of the reverse embedding of $\PS$ is as follows:
1354: first, we take the original conditional expression
1355: $e=e((a_1|b_1),\dots,(a_m|b_m))$ and replace every $(a_i|b_i)$ occurring
1356: in it by $b_i^\C\Since (a_i\land b_i),$ obtaining $e'\in \TL,$ and then define
1357: $\ttr{e}:=(e'|\TRUE).$ Formally:
1358:
1359:
1360: \begin{equation}\label{drugi}
1361: \begin{split}
1362: \ttlr{(a| b)}&=b^\C\Since(a\land b)\\
1363: \ttlr{e \land e'}&=\ttlr{e}\land\ttlr{e'}\\
1364: \ttlr{e \lor e'}&=\ttlr{e}\lor \ttlr{e'}\\
1365: \ttlr{\llnot e}&=\lnot\ttlr{e}\\
1366: \ttr{e}&=(\ttlr{e}|\TRUE).
1367: \end{split}
1368: \end{equation}
1369:
1370:
1371: For this particular embedding, we have the following consequence of
1372: Theorem \ref{reverse} (and the simple fact that reversing is an
1373: automorphism of the whole $\TT$)
1374:
1375: \begin{theorem}\label{dwie-zaby} $\ttr{\cdot}$ is an embedding of
1376: the $\PS$ \cea{} into $\CC,$ in the sense that for any underlying
1377: probability space, and any conditional expressions $e,e'$,
1378: $\ttr{e}=\ttr{e'}$ iff $\ps{e}=\ps{e'},$ and
1379: $\Pr(\ttr{e})=\Pr_o(e).$\qed
1380: \end{theorem}
1381:
1382:
1383: \subsection{Sparse reverse embedding}
1384:
1385: We give a new, radically different interpretation of $\PS$ in $\CC.$ The
1386: main difference is that it is not an embedding. We are going to present
1387: a way to interpret $\PS$ expressions in $\CC$ so that, for any
1388: probability space $(\Omega,\PP(\Omega),\Pr),$ the $\Pr_o$-probability of
1389: a $\PS$-expression is equal to the asymptotic probability of its
1390: interpretation. However, expressions which yield equal element of the
1391: $\PS$ \cea, may well give distinct conditional events in $\CC,$
1392: although, as said before, these expressions will have equal asymptotic
1393: probabilities.
1394:
1395: First of all , we redefine the meaning of simple conditionals $(a|b).$
1396: According to the new embedding, they are represented by $\TT$ formula
1397:
1398: \[(a\land\lnot(\PBOX\lnot b\,\lor\,\lnot\PREV\TRUE)|b\lor \lnot\PDIA
1399: b).\]
1400:
1401: Denote this formula by $(a|_Sb).$
1402:
1403: It is essentially the simple conditional $(a|b)\in\TT,$ except that it
1404: is defined and false until $b$ becomes true for the very first time, and
1405: since then behaves exactly like $(a|b)$ does in $\TT$. The manipulation
1406: is necessary to accommodate the $\PS$ principle, that degenerate simple
1407: conditionals, like $(a|0),$ do have probability, and that it is $0.$
1408:
1409:
1410: We now define the sparse reverse interpretation of $\PS$ in $\CC$ as
1411: follows: First, we take the original conditional expression
1412: $e=e((a_1|b_1),\dots,(a_m|b_m))$ and set
1413: $\ttrs{e}=(\ttlr{e}|\Lor_{i=1}^m(b_i\lor \lnot\PDIA b_i)),$ where
1414: $\ttlr{\cdot}$ has been defined in \eqref{drugi}.
1415:
1416: The conjunction and disjunction are defined precisely when at least
1417: one of the arguments is defined, so they resemble the connectives of
1418: $\SAC$ in this respect, but instead of assigning the other arguments
1419: default values when they are undefined, like $\SAC$ does, their most
1420: recent defined values are always used, instead. Here is an example
1421: Moore machine, in which we use $\epsilon$-moves. The graphical
1422: representation in Fig. \ref{f00} and Fig. \ref{f0} shows that the new
1423: operation is closely related to the reversed product space
1424: conjunction, as can be expected from the shape of the $\TT$
1425: representation.
1426:
1427:
1428: \begin{figure}[htp]
1429:
1430: \[
1431: \xymatrix @C=16mm @R=36mm
1432: {
1433: &&*+++[o][F]{0_{1\bot}}
1434: \ar@(ur,ul)[]
1435: \ar[r]_{a^\C bd^\C}
1436: &*+++[o][F]{0_{0\bot}}
1437: \ar@(ur,ul)[]
1438: \ar@<-2ex>[l]_{abd^\C}
1439: \\
1440: {}\ar[r]&*+++[o][F]{0_{\bot\bot}}
1441: \ar@(ul,u)[]
1442: \ar[ur]|{abd^\C}
1443: \ar[urr]|{a^\C bd^\C}
1444: \ar[dr]|{b^\C cd}
1445: \ar[drr]|{b^\C b^\C d}
1446: \\
1447: &&*+++[o][F]{0_{\bot1}}
1448: \ar@(dr,dl)[]
1449: \ar[r]^{b^\C c^\C d}
1450: &*+++[o][F]{0_{\bot0}}
1451: \ar@(dr,dl)[]
1452: \ar@<2ex>[l]^{b^\C cd}
1453: }
1454: \]
1455:
1456: \caption[Sparse reverse conjunction 1.]{Moore machine of
1457: $\ttrs{(a|b)\land(c|d)},$ part 1. This part of the machine is the
1458: transient part of the Markov chain, when $\Pr(b),\Pr(d)>0$ (Lemma
1459: \ref{12}) and the whole reachable part when at least one of these
1460: probabilities is $0$ (Lemma \ref{11}).\\ Subscripts of the state
1461: labels indicate the most recent value of $(a|b)$ and $(c|d),$
1462: respectively.}\label{f00}
1463: \end{figure}
1464:
1465:
1466: \begin{figure}[htp]
1467:
1468: \[\UseTips
1469: \xymatrix @C=16mm @R=36mm
1470: {
1471: {}\save[]+<0cm,-18mm>*+++[o][F-]{\bot}
1472: \ar@/^/[d]^{\epsilon}
1473: \ar@/_/@{<-}[d]_{b^\C d^\C}
1474: \restore
1475: &&*+++[o][F-]{0_{01}}
1476: \ar@(ul,ur)
1477: \ar@<2ex>[dll]|(.36){c^\C d\land (ab)^\C}
1478: \ar@<2ex>[ddl]|(.36){abc^\C d}
1479: \ar@<2ex>[dr]|(.36){ab\land (c^\C d)^\C}
1480: \ar@/^/[r]^{b^\C d^\C}
1481: &*+++[o][F-]{\bot}
1482: \ar@/^/[l]^{\epsilon}
1483: \\
1484: *+++[o][F-]{0_{00}}
1485: \ar@(l,ul)
1486: \ar@<2ex>[rrr]|(.36){abcd}
1487: \ar@<2ex>[rru]|(.36){cd\land (ab)^\C}
1488: \ar@<2ex>[dr]|(.36){ab\land (cd)^\C}
1489: &&&*+++[o][F-]{1_{11}}
1490: \ar@(ur,r)
1491: \ar@<2ex>[lll]|(.36){a^\C bc^\C d}
1492: \ar@<2ex>[lld]|(.36){c^\C d\land (a^\C b)^\C}
1493: \ar@<2ex>[lu]|(.36){a^\C b\land (c^\C d)^\C}
1494: \\
1495: *+++[o][F-]{\bot}
1496: \ar@/_/[r]_{\epsilon}
1497: %\save[]+<-2cm,3cm>\ar[u]\restore %strzalka rozpoczynajaca rysunek
1498: &*+++[o][F-]{0_{10}}
1499: \ar@(rd,dl)
1500: \ar@<2ex>[rru]|(.36){cd\land(a^\C b)^\C}
1501: \ar@<2ex>[uur]|(.36){a^\C bcd}
1502: \ar@<2ex>[ul]|(.36){a^\C b\land (ab)^\C}
1503: \ar@/_/[l]_{b^\C d^\C} &
1504: &{}\save[]+<0cm,18mm>*+++[o][F-]{\bot}
1505: \ar@/^/[u]^{\epsilon}
1506: \ar@/_/@{<-}[u]_{b^\C d^\C}
1507: \restore
1508: }
1509: \]
1510:
1511: \caption[Sparse reverse conjunction 2.]{Moore machine of $\ttrs{(a|b)\land(c|d)},$ part 2. This part
1512: of the machine is the (only) ergodic class of the Markov chain, when
1513: $\Pr(b),\Pr(d)>0$ (Lemma \ref{12}), and is unreachable when at
1514: least one of these probabilities is $0$ (Lemma \ref{11}).\\ Subscripts
1515: of the state labels indicate the most recent value of $(a|b)$ and
1516: $(c|d),$ respectively.} \label{f0}
1517: \end{figure}
1518:
1519: \begin{figure}[htp]
1520: \[
1521: \begin{array}{|c|c|c|}
1522: \hline
1523: \multicolumn{3}{|c|}{\text{Missing arrows}}\\
1524: \hline\hline
1525: \text{From}&\text{To}&\text{Label}\\
1526: \hline
1527: 0_{\bot\bot},0_{\bot0},0_{\bot1},0_{1\bot},0_{0\bot}&0_{00}&a^\C bc^\C d
1528: \\
1529: \hline
1530: 0_{\bot\bot},0_{\bot0},0_{\bot1},0_{1\bot},0_{0\bot}&0_{01}&a^\C bcd
1531: \\
1532: \hline
1533: 0_{\bot\bot},0_{\bot0},0_{\bot1},0_{1\bot},0_{0\bot}&0_{10}&abc^\C d
1534: \\
1535: \hline
1536: 0_{\bot\bot},0_{\bot0},0_{\bot1},0_{1\bot},0_{0\bot}&1_{11}&abcd
1537: \\
1538: \hline
1539: 0_{1\bot}&1_{11}&b^\C cd
1540: \\
1541: \hline
1542: 0_{0\bot}&0_{01}&b^\C cd
1543: \\
1544: \hline
1545: 0_{1\bot}&0_{10}&b^\C c^\C d
1546: \\
1547: \hline
1548: 0_{0\bot}&0_{00}&b^\C c^\C d
1549: \\
1550: \hline
1551: 0_{\bot1}&1_{11}&abd^\C
1552: \\
1553: \hline
1554: 0_{\bot0}&0_{10}&abd^\C
1555: \\
1556: \hline
1557: 0_{\bot1}&0_{01}&a^\C bd^\C
1558: \\
1559: \hline
1560: 0_{\bot0}&0_{00}&a^\C bd^\C\\
1561: \hline
1562: \end{array}
1563: \]
1564: \caption[Sparse reverse conjunction 3.]{Moore machine of $\ttrs{(a|b)\land(c|d)},$ part 3. This part
1565: of the machine is the table of the transitions from the ``transient''
1566: part (Fig.~\ref{f00}) to the ``ergodic'' part
1567: (Fig.~\ref{f0}).}\label{f000}
1568: \end{figure}
1569:
1570:
1571: \begin{lemma}\label{11}
1572: If $\Pr(b_i)=0$ for at least one $1\leq i\leq m,$ then
1573: $\Pr(\ttrs{e})=\Pr(\ttt{e}).$
1574: \end{lemma}
1575: \begin{proof} In this case, assuming $\Pr(b_i)=0$, we have that
1576: $\Lor_{i=1}^m(b_i\lor \lnot\PDIA b_i))$ is true with probability $1,$
1577: hence $\ce{\ttrs{e}}=\ce{\ttr{e}}$ with probability $1.$
1578: Now the thesis follows immediately from Theorem \ref{dwie-zaby}.
1579: \end{proof}
1580:
1581:
1582: \begin{lemma}\label{12} If $\Pr(b_i)>0$ for all $1\leq i\leq m,$ then
1583: $\Pr(\ttrs{e})=\Pr_o(\ttt{e}).$
1584: \end{lemma}
1585: \begin{proof} Let $e=e((a_1|b_1,\dots,(a_n|b_n)).$ We prove
1586: $\Pr(\ttr{e})=\Pr(\ttrs{e}),$ which is, by Theorem
1587: \ref{dwie-zaby}, equivalent to what we have to show.
1588:
1589: Denote $t=$ the first moment $m$ when all the $b_i$'s have already
1590: been defined.
1591:
1592: Let us note that, if $t<n,$ then the event
1593: $\ce{\ttrs{e}}_n=\bot$ is independent of the whole history of
1594: $\ce{\ttrs{e}}$ up to time $n-1.$ This is so because the decision
1595: whether $\ce{\ttrs{e}}_n$ is defined or not depends solely on the
1596: present time values of $b_i$'s, and their present time values become
1597: independent of the (strict) past, when the condition
1598: $\Lor_{i=1}^n\lnot\PDIA b_i$ becomes for the first time false,
1599: because it remains then false forever, and the ``given'' part of
1600: $\ttrs{e}$ does not contain any other time modalities.
1601:
1602:
1603: Moreover, whenever
1604: $\ce{\ttrs{e}}(\omega_1\dots\omega_n)\neq\bot,$ then in fact
1605: $\ttrs{e}(\omega_1\dots\omega_n)=
1606: \ttr{e}(\omega_1\dots\omega_n),$ which is clear from the
1607: syntactic representation of both conditional objects.
1608:
1609:
1610:
1611: Denote for convenience
1612:
1613: $q=\Pr(\ce{\ttrs{e}}_n=\bot|t<n)=1-\Pr(\Lor_{i=1}^m b_i)<1,$
1614: as well as and $c=\ttr{e}$ and $c^\S=\ttrs{e}.$
1615:
1616: Fix $\varepsilon>0.$ Let $M$ be large enough to have $\Pr(t\geq
1617: M)<\varepsilon.$ Let $N$ be a large integer, and let $n$ satisfy
1618: $n-N>M.$
1619:
1620: We have then by the independence
1621:
1622: \begin{align*}
1623: \Pr(\ce{c}_n=1)&\geq
1624: \sum_{i=0}^N\Pr(\ce{c^\S}_{n-i}=1)\Pr(\ce{c^\S}_{n-i+1}=\bot)
1625: \dots \Pr(\ce{c^\S}_{n}=\bot)-\varepsilon,\\
1626: \Pr(\ce{c}_n=0)&\geq
1627: \sum_{i=0}^N\Pr(\ce{c^\S}_{n-i}=0)\Pr(\ce{c^\S}_{n-i+1}=\bot)
1628: \dots \Pr(\ce{c^\S}_{n}=\bot)-\varepsilon,
1629: \end{align*}
1630:
1631: where the $\varepsilon$ error terms are caused by the event that $t\geq M.$
1632:
1633: Because each of the $\Pr(\dots)$ expressions above tends to a limit as
1634: $n$ approaches infinity, we get
1635:
1636: \begin{align*}
1637: \Pr(c)&\geq
1638: \sum_{i=0}^N\lim_{n\to\infty}\Pr(\ce{c^\S}_{n-i}=1)\lim_{n\to\infty}\Pr(\ce{c^\S}_{n-i+1}=\bot)
1639: \cdots \Pr(\ce{c^\S}_{n}=\bot)-\varepsilon,\\
1640: &=\sum_{i=0}^N\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=\bot)
1641: \cdots \Pr(\ce{c^\S}_{n}=\bot)-\varepsilon\\
1642: &=\sum_{i=0}^N\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)(\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=\bot))^i-\varepsilon\\
1643: &=\sum_{i=0}^N\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)q^i-\varepsilon\\
1644: &=\frac{1-q^N}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)-\varepsilon,
1645: \end{align*}
1646:
1647: and similarly
1648:
1649: \[1-\Pr(c)\geq
1650: \frac{1-q^N}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=0)-\varepsilon.
1651: \]
1652:
1653: In the limit $N\to\infty$ both inequalities become
1654:
1655: \begin{align*}
1656: \Pr(c)&\geq
1657: \frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)-\varepsilon\\
1658: 1-\Pr(c)&\geq
1659: \frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=0)-\varepsilon,
1660: \end{align*}
1661:
1662: hence
1663:
1664: \begin{align*}
1665: \Pr(c)&\geq
1666: \frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)-\varepsilon\\
1667: \Pr(c)&\leq
1668: 1-\frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=0)+\varepsilon\\
1669: &=\frac{1}{1-q}(1-q -\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=0))+\varepsilon\\
1670: &=\frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1)+\varepsilon,
1671: \end{align*}
1672:
1673: because
1674: $1=\Pr(\ce{c^\S}_{n}=1)+\Pr(\ce{c^\S}_{n}=0)+\Pr(\ce{c^\S}_{n}=\bot),$
1675: in which the last term is constant equal $q.$
1676:
1677: We took an arbitrary $\varepsilon>0,$ therefore indeed
1678:
1679: \[
1680: \Pr(c^\S)=\frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=1),
1681: \]
1682:
1683: and likewise
1684:
1685: \[
1686: 1-\Pr(c^\S)=\frac{1}{1-q}\lim_{n\to\infty}\Pr(\ce{c^\S}_{n}=0).
1687: \]
1688:
1689: {}From the last two equalities the equality $\Pr(c)=\Pr(c^\S)$
1690: follows immediately.
1691: \end{proof}
1692:
1693: Summing up,
1694:
1695: \begin{theorem}\label{sparse} For every conditional expression $e$ and
1696: every probability assignment to the elements in $\Omega,$
1697: $\Pr_o(e)=\Pr(\ttrs{e}).$\qed
1698: \end{theorem}
1699:
1700: The very important consequence of the theorem is the following:
1701:
1702: \begin{corollary} The formalism of $\PS$ \cea{}, seen as a logic of
1703: conditionals, is unable to determine certain probabilistic
1704: characteristics, other than the asymptotic probability, associated with
1705: stochastic processes.
1706: \end{corollary}
1707: \begin{proof} $\PS$ possesses two interpretations in $\CC,$ one of which
1708: consists entirely of always defined temporal conditionals (the
1709: $\ttt{\cdot}$ embedding), while the second contains conditionals which
1710: are defined with asymptotic probability strictly less than 1 (the
1711: $\ttrs{\cdot}$ interpretation).\end{proof}
1712:
1713: Another similar example of deficiencies of the $\PS$ \cea{} can be found
1714: below, Theorem \ref{(a|0)}.
1715:
1716: Let us note that the $\ttrs{\cdot}$ interpretation does not preserve
1717: the algebraic structure of the $\PS$ \cea, in general. Indeed, already
1718: $\ps{(0|a)}=\ps{(0|b)}$, while $\ttrs{(0|a)}=(a|_Sb)$ and
1719: $\ttrs{(0|b)}=(0|_Sb)$ represent different conditional objects, for
1720: $a\neq b,$ $a,b\in\E.$
1721:
1722: \section{Advantages of $\TT$ conditionals}
1723:
1724: \subsection{Complexity and proof systems}
1725:
1726: The paper \cite{g94} asks for the proof systems for various
1727: three-valued logics, appearing in the context of the theory of
1728: conditionals.
1729:
1730: In order to discuss this issue, we use the machinery of complexity
1731: theory. All the necessary definitions can be found in \cite{HU}.
1732:
1733: As we prove below, for all of the major \cea's, the sets of weak
1734: tautologies are $\coNP$ complete.
1735:
1736: In the light of the above results, there is a little hope for a
1737: practically useful proof system for the most prominent systems among
1738: $\GNW$ and $\SAC.$ Indeed, unless $\NP=\coNP,$ a very unlikely
1739: complexity-theoretic collapse, for {\em every\/} sound and complete
1740: proof system for the two above logics, there must be weak tautologies of
1741: length $n$ such that their shortest proofs are of superpolynomial length
1742: w.r.t.\ $n,$ for infinitely many $n.$
1743:
1744: Temporal logic is known to be $\PSPACE$-complete, Theorem \ref{Emerson}.
1745: It is therefore obvious that the set of weak tautologies of $\TT,$
1746: consisting of all expressions $(\varphi|\psi)$ such that for every
1747: $w\in\trzy^+,$ $(\varphi|\psi)(w)\in\{1,\bot\}$ (equivalently: that
1748: $\psi\to\varphi$ is a tautology of $\TL$), is $\PSPACE$-complete, too.
1749:
1750: Although it is commonly believed that $\coNP\subsetneq\PSPACE,$ from
1751: practical standpoint both admit exponential time algorithms, and no
1752: better ones are known. Consequently, the practical algorithmic
1753: difference between \cea's and $\TT$ is not so crucial. The advantage of
1754: considering \cea's as subsystems of $\TT$ stems from the fact that a lot
1755: is known about proof systems for temporal logic --- unlike for \cea's.
1756:
1757:
1758: We are not interested in the complexity of \cea's as logics involving
1759: terms $(a|b)$ as atoms, but rather as $\trzy$-valued logics. To explain
1760: the difference, let us note that in the calculus of any of the \cea's
1761: one can easily restrict atoms to be two-valued (e.g., by using only
1762: atoms of the form $(a|1)$), and thus all the complexity questions
1763: trivialize, the sets of tautologies of all the logics are
1764: $\coNP$-complete. Here, we assume that the atoms can always assume all
1765: three logical values, and are effectively variables. Hence we indeed
1766: view our (weak) tautologies as a kind of (weak) meta-tautologies, i.e.,
1767: formulas which evaluate to either $1$ or $\bot,$ no matter what the
1768: arguments are.
1769:
1770: Formally, for this section we modify $\LL$ and $\LL^|$ replacing simple
1771: conditionals $(a|b)$ by variables $p_1,p_2,\dots$ in conditional
1772: expressions.
1773:
1774: Likewise, given a valuation $v:\{p_1,p_2,\dots\}\to\trzy,$ we let
1775: $\sac{\cdot}^v$ and $\gnw{\cdot}^v$ assign values in $\trzy$ to
1776: expressions in $\LL^|$. These values are determined by the equations in
1777: \eqref{sac} and \eqref{gnw}.
1778:
1779: An expression $e$ is a {\em weak tautology\/} of $\SAC$ \cea{} ($\GNW$
1780: \cea, respectively) iff $\sac{e}^v\neq 0$ ($\gnw{e}^v\neq 0,$
1781: respectively) for every $v$.
1782:
1783: $e$ is a strong tautology of $\SAC$ ($\GNW$, respectively), iff the
1784: above values are 1 for every $v.$
1785:
1786: We consider the complexity problem of determining if an expression $e$
1787: is a weak tautology according to each of the considered \cea's,
1788: considering also some syntactical restrictions put on the syntactical
1789: shape of $e.$
1790:
1791: We do not consider strong tautologies, which is explained by the
1792: following.
1793:
1794: \begin{proposition} In $\GNW$\/ and $\SAC$\/ there are no strong
1795: tautologies.
1796: \end{proposition}
1797: \begin{proof} All connectives have value $\bot$ if all their arguments
1798: are $\bot,$ for both $\SAC$ and $\GNW.$
1799: \end{proof}
1800:
1801: Occasionally, we want to consider $\LL^|$ as the syntax of classical
1802: logic. In this case, given a valuation $v:\{p_1,p_2,\dots\}\to\dwa,$
1803: the value $\cl{e}^v\in\dwa$ is computed according to the classical
1804: rules, where the conditioning $|$ is understood as reverse implication:
1805: $(a|b)$ is $a\leftarrow b.$
1806:
1807: \paragraph{Pure conditional parts.}
1808:
1809: We consider here pure conditional fragments of $\SAC$ and $\GNW,$ i.e.,
1810: expressions in which the only connective used is $|.$
1811:
1812: It shows that unlimited use of re-conditioning leads to
1813: $\coNP$-completeness of the weak tautology problem.
1814:
1815: \begin{theorem}\label{pure|} It is an $\coNP$-complete problem to
1816: determine if an expression $e\in\LL^|$ involving only re-conditioning is
1817: a weak tautology of $\SAC.$
1818:
1819: It is an $\coNP$-complete problem to determine if an expression
1820: $e\in\LL^|$ involving only re-conditioning is a weak tautology of
1821: $\GNW.$
1822: \end{theorem}
1823: \begin{proof}
1824: It is obvious that the sets of weak tautologies in both cases are in
1825: $\coNP.$ So it remains to prove their hardness in this complexity class.
1826:
1827: It is easily seen that the (re-)conditioning operators of both $\GNW$
1828: and $\SAC$ satisfy the following property: the equivalence relation
1829: $\approx$ on $\trzy $ identifying $1$ with $\bot$ is a congruence of
1830: $\A=\langle\trzy ,{|_\SAC}\rangle$ and $\B=\langle\trzy
1831: ,{|_\GNW}\rangle,$ and the quotient algebras both $\A=\langle\trzy
1832: ,{|_\SAC}\rangle/{\approx}$ and $\B=\langle\trzy
1833: ,{|_\GNW}\rangle/{\approx}$ are isomorphic to the 2-element algebra with
1834: the reversed classical implication $\langle\dwa ,\leftarrow\rangle.$ The
1835: natural epimorphism $\eta_\A:\A\to\langle\dwa ,\leftarrow\rangle$ sends 1 and
1836: $\bot$ to 1, and 0 to 0, and the definition of $\eta_\B$ is identical.
1837:
1838: Therefore a pure conditional expression is a weak tautology of either of
1839: the considered \cea's iff it is a classical tautology, after its
1840: (re-)conditioning operator is replaced by the reversed classical
1841: implication. The classical formula resulting from this replacement is
1842: denoted $\bar{e}.$
1843:
1844: We have to prove that $e$ is not a weak tautology iff $\bar{e}$
1845: is not a tautology. Let $v$ be any valuation of the variables of
1846: $e$ in $\trzy .$ Now we use the natural epimorphism $\eta_\A$
1847: and get
1848:
1849: \[\cl{\bar{e}}^{\eta_\A\circ v}=\eta_\A(\sac{e}^v).\]
1850:
1851: So if one of the values above can be $0,$ the other can be, as well,
1852: which establishes the desired equivalence.
1853:
1854: Since it is known that the tautologies of the classical propositional
1855: logic of pure implication are $\coNP$ complete \cite{H}, the claim
1856: follows.
1857: \end{proof}
1858:
1859: As a by-product we have
1860:
1861: \begin{corollary} The sets of pure conditional weak tautologies of
1862: $\SAC$\/ and $\GNW$\/ are identical.\qed
1863: \end{corollary}
1864:
1865:
1866: \paragraph{Flat parts.}
1867:
1868: Here we consider $\SAC$ and $\GNW$ without re-conditioning.
1869:
1870: We can define the following $\NP$-complete problem 3CNF-SAT.
1871:
1872: Given: an expression $e\in \LL$ of the following syntactical form:
1873:
1874: \begin{equation}\label{3CNF-SAT}
1875: e=(\ell_{11}\lor \ell_{12}\lor \ell_{13})\land (\ell_{21}\lor
1876: \ell_{22}\lor \ell_{23})\land \dots\land (\ell_{s1}\lor \ell_{s2}\lor
1877: \ell_{s3}),\end{equation}
1878:
1879: where each of the $\ell_{ij}$ is either $p_j$ or $\llnot p_j.$
1880:
1881: The $\NP$-complete problem is: given $e$ of the above shape, determine
1882: if $e$ is satisfiable, i.e, if there exists $v$ such that $\cl{e}^v=1.$
1883:
1884: It follows that it is $\coNP$-complete to determine, given $e$ as above,
1885: if $e$ is {\em not\/} satisfiable, i.e., whether $\cl{\llnot e}^v=1$ for
1886: every $v.$
1887:
1888: In order to prove $\coNP$-completeness of the sets of weak tautologies
1889: of either of the \cea's, we have to construct a polynomial time
1890: computable transformation $e\mapsto \bar{e}$ translating $e$ of the form
1891: \eqref{3CNF-SAT} into $\bar{e}$ of the form conforming to the restriction set in
1892: the respective theorem, and such that $\llnot e$ is not satisfiable in the
1893: classical sense iff $\bar{e}$ is a weak tautology of the respective
1894: logic.
1895:
1896:
1897:
1898: \begin{theorem} It is an $\coNP$-complete problem to determine if an
1899: expression $e\in\LL$ is a weak tautology of $\SAC.$
1900:
1901: It is an $\coNP$-complete problem to determine if an expression
1902: $e\in\LL$ is a weak tautology of $\GNW.$
1903: \end{theorem}
1904:
1905: \begin{proof} It is easily seen that the connectives $\land_\GNW$ and
1906: $\lor_\GNW$ satisfy again the property that the equivalence relation
1907: $\approx$ on $\trzy$ identifying $1$ with $\bot$ is a congruence of the
1908: algebra with the above functions, and the quotient algebra
1909: $\langle\trzy, \land_\GNW,\lor_\GNW\rangle\big/\approx$ is isomorphic to
1910: the classical $\langle\dwa,\land,\lor\rangle.$ This fails about the
1911: negation, however.
1912:
1913: As the negation is applied to atoms only in 3CNF-SAT, we do not have to use
1914: the negation of $\GNW$ directly. Instead, we introduce new variables to
1915: denote the negations, and force them to behave correctly outside of the
1916: translation of $e.$
1917:
1918: Formally, let the mapping $e\mapsto e'$ from the classical propositional
1919: logic into $\GNW$ be defined by replacing unnegated atoms $p$ in $e$ by
1920: $\hat{p}$ and negated atoms $\llnot p$ by $\check{p}.$ Concerning binary
1921: connectives, we leave $\land$ and $\lor$ untouched.
1922:
1923: Then let $\bar{e}$ be defined as $\llnot(e'\land{\Land\limits_p}
1924: (\hat{p}\lor \check{p})\land(\llnot\hat{p}\lor\llnot \check{p})),$ where
1925: $p$ in the big conjunction ranges over all propositional variables of
1926: $e.$
1927:
1928: Certainly the mapping $e\mapsto\bar{e}$ is computable in polynomial
1929: time. In order to show the $\coNP$ completeness of the set of
1930: tautologies of $\GNW,$ it suffices to show two implications:
1931:
1932: \begin{itemize}
1933: \item if $e$ is satisfiable classically, then $\bar{e}$ is not a weak
1934: tautology of $\GNW.$
1935: \item if then $\bar{e}$ is not a weak tautology of $\GNW,$ then $e$ is
1936: satisfiable classically.
1937: \end{itemize}
1938:
1939: For the first item, assume that $e$ is satisfiable, i.e., there is an
1940: assignment $v$ of $0$'s and $1$'s to the propositional variables of $e$
1941: which makes $e$ into $1.$ We construct a $\trzy$-valued assignment $w$
1942: which makes $\bar{e}$ into $0$. If $v(p)=1,$ we let $w(\hat{p})=1$ and
1943: $w(\check{p})=0.$ If $v(p)=0,$ we let $w(\hat{p})=0$ and
1944: $w(\check{p})=1.$ In $e'$ each variable has under $w$ exactly the value
1945: of the corresponding literal in $e$ has under $v.$ So $e'$ evaluates to
1946: $1,$ because connectives in $\GNW$ behave classically for classical
1947: arguments. In addition, each of the formulas $(\hat{p}\lor
1948: \check{p})\land(\llnot\hat{p}\lor\llnot\check{p})$ evaluates to $1,$ so
1949: altogether $\bar{e}$ evaluates to the $\llnot_0$-negation of the value
1950: to which $e'$ does evaluate, which is $0,$ as desired.
1951:
1952: For the second item, assume there is an assignment $w$ of $0$'s, $1$'s
1953: and $\bot$'s to the propositional variables of $\bar{e}$ which makes it
1954: $0.$ It follows that each of the terms
1955: $(\hat{p}\lor\check{p})\land(\llnot\hat{p}\lor\llnot\check{p})$ must
1956: evaluate to $1$ under $w.$ Therefore of each pair $\hat{p},\check{p},$
1957: one variable must be assigned $1$ and the other $0$ by $w,$ which can be
1958: checked by simple inspection of all possibilities. Moreover, $e'$ must
1959: evaluate to $1$ under $w,$ which is indeed $\dwa$-valued, by the
1960: previous observation. The connectives of $\GNW$ act classically for
1961: classical arguments, therefore $e$ is indeed classically satisfiable, by
1962: the valuation $v:p\mapsto w(\hat{p}).$
1963:
1964: This finishes the proof.
1965: \end{proof}
1966:
1967:
1968: \begin{theorem} The weak flat-conditional $\SAC$ is $\coNP$-complete.
1969: \end{theorem}
1970: \begin{proof} We are going to use the same proof idea as before.
1971: However, we have a small problem. The conjunction of $\SAC$ does not
1972: permit us to deduce, that if a conjunction of two formulas evaluates to
1973: $1,$ so does each of the components.
1974:
1975: So instead of the original conjunction, we have to use some custom
1976: connective defined from the conjunction, disjunction and negation, which
1977: will act as a ``good'' conjunction, for which the inference does hold.
1978: It turn out, that the conjunction of $\GNW$ is not definable in $\SAC,$
1979: but there is another connective we can use instead, and which is
1980: definable (we discuss the definability of connectives in $\SAC$ and
1981: $\GNW$ in another paper \cite{SAC-GNW}). Its definition is as follows:
1982: \[x\sqcap y\equiv [ x\lor( y\land( x\lor\llnot y))]\land[ y\lor( x\land
1983: (y\lor\llnot x))].\]
1984:
1985: It is not difficult (but tedious) to check, that $\sac{x\sqcap y}$ has
1986: truth table
1987:
1988: \[\conn\sqcap00001000,\]
1989:
1990:
1991: which is exactly what we need for our purposes. The only subtle point
1992: is that our $x\sqcup y$ is substantially longer than $|x|+|y|.$ Indeed
1993: it is about 4 times longer. We do replace $\land$ by $\sqcap$ in very
1994: long conjunctions. However, if we represent this long conjunction as a
1995: balanced binary (parse) tree, i.e., insert brackets to obtain the
1996: structure
1997: \[(((\ldots\sqcap\ldots)\sqcap(\ldots\sqcap\ldots))\sqcap((\ldots\sqcap\ldots)\sqcap(\ldots\sqcap\ldots))),\]
1998: the depth of nesting of conjunctions is at most log base 2 of the number
1999: $N$ of clauses in the conjunction, and the total increase of length
2000: caused by the replacement is $4^{\text{depth of
2001: nesting}}=4^{\log_2N}=N^2.$ Altogether, the resulting formula, using
2002: $\sqcap$ in place of $\land,$ is still of polynomial size, and can be
2003: easily constructed in polynomial time, as needed.
2004: \end{proof}
2005:
2006:
2007: \subsection{Independence of conditional events}
2008:
2009: There has been a considerable amount of interest in the independence
2010: issue for conditional events, reflected in the \cea{} literature
2011: \cite{g94,c97,p88}. The problem is that typically even for $a$ and $b$
2012: mutually independent of $c$ and $d$ one does not have
2013: $\Pr((a|b)\land(c|d))=\Pr((a|b))\Pr((c|d)).$ The only exception is
2014: $\PS,$ where this equality holds. The other variant of independence:
2015: $\Pr((a|b)|(c|d))=\Pr((a|b))$ is undefined in some formalisms, due to
2016: the lack of re-conditioning operator, and fails in others. However,
2017: note that in the \cea{} framework one cannot obtain any proper
2018: characterization of independence, because there is no underlying
2019: probabilistic semantics, in which one could say which pairs of
2020: conditionals are independent and which aren't, and then attempt to
2021: characterize this by equalities among probabilities. One {\em feels\/}
2022: that $(a|b)$ and $(c|d)$ should be independent for mutually independent
2023: arguments, but this is not more than a feeling, and there is no idea
2024: there what might make two conditionals independent when their arguments
2025: are not mutually independent, or when they are composite.
2026:
2027: We can address this problem in our semantical setting. First of all,
2028: for $a$ and $b$ mutually independent of $c$ and $d,$ the stochastic
2029: processes $\ce{(a|b)}$ and $\ce{(c|d)}$ are obviously independent. And
2030: of course, the {\em independence of the stochastic processes\/} is what
2031: the independence of conditionals should be. This remains true, no matter
2032: which \cea\ we consider. It is, however, a different story if this
2033: independence can be formally characterized in terms of equalities
2034: between probabilities of conditionals in the \cea\ under consideration.
2035: It appears that in the pure \cea{} formalism this cannot be achieved,
2036: because in Theorem \ref{PS-indep} below we show that independence is
2037: undefinable in the $\PS$ \cea.
2038:
2039: To be precise, the full independence of stochastic processes $\X,\Y$
2040: means that {\em the full histories of both processes\/} are
2041: independent, which is different from the much less restrictive requirement that
2042: just the present time values should be independent. The first version
2043: is formalized by the requirement that $\X_{+,t}$ and $\Y_{+,t}$ are
2044: independent at any time $t>0,$ i.e., for any $w_1\dots w_t,v_1\dots
2045: v_t\in\trzy^t$ holds
2046:
2047: \begin{multline*}
2048: \Pr\left(\hspace{-6pt}
2049: \begin{array}{ccc}
2050: X_1=w_1,&\dots,&X_t=w_t\\
2051: Y_1=v_1,&\dots,&Y_t=v_t
2052: \end{array}
2053: \hspace{-6pt}\right)=\\
2054: \Pr(X_1=w_1,\dots,X_t=w_t)\Pr(Y_1=v_1,\dots,Y_t=v_t).
2055: \end{multline*}
2056:
2057: The weaker, present tense independence requires only that $X_t$ and $Y_{t}$
2058: are independent at any time $t>0,$ i.e., that for any $w,v\in\trzy$ holds
2059: $\Pr(X_t=w,Y_t=v)=\Pr(X_t=w)\Pr(Y_t=v).$ To see the difference it is worth
2060: noting that for {\em any present tense\/} $\TT$ formula $(\varphi|\psi)$ the
2061: processes $\ce{(\varphi|\psi )}$ and $\ce{(\PREV \varphi|\PREV \psi )}$ are
2062: present tense independent, although of course they are easily seen to be
2063: dependent, unless the former is constant.
2064:
2065: But let us note the following simple fact.
2066:
2067: \begin{lemma}\label{present-tense} If $c_1$ and $c_2$ are two present
2068: tense conditionals, they are independent iff they are present tense
2069: independent.\qed
2070: \end{lemma}
2071:
2072: We know now what independence should {\em mean.} It is another story how to
2073: {\em characterize\/} it in terms of the asymptotic probability of
2074: conditionals.
2075:
2076: First we prove the characterization for present tense independence at fixed
2077: time.
2078:
2079: Let $\defn (a|b):=(b|\TRUE).$
2080:
2081: \begin{lemma}\label{indep}
2082: Let $n$ be a fixed time instant. The following are equivalent:
2083: \begin{itemize}
2084: \item Random variables $\ce{(a|b)}_n$ and $\ce{(c|d)}_n$ are
2085: independent.
2086: \item The following four equalities hold:
2087: \begin{align}
2088: \pr_n((a|b)\land_\Sch(c|d))&=\pr_n((a|b))\pr_n((c|d))\label{i1}\\
2089: \pr_n((a|b)\land_\Sch\defn(c|d))&=\pr_n((a|b))\pr_n(\defn(c|d))\label{i2}\\
2090: \pr_n(\defn(a|b)\land_\Sch(c|d))&=\pr_n(\defn(a|b))\pr_n((c|d))\label{i3}\\
2091: \pr_n(\defn(a|b)\land\defn(c|d))&=\pr_n(\defn(a|b))\pr_n(\defn(c|d))\label{i4},
2092: \end{align}
2093: where we assume an equation to hold in case when both sides are
2094: undefined.
2095: \end{itemize}
2096: \end{lemma}
2097: \begin{proof}
2098: $\Downarrow$ Independence of random variables $\ce{(a|b)}_n$ and
2099: $\ce{(c|d)}_n$ implies, in particular, that
2100:
2101: \begin{equation}\tag{$\ref{i4}'$}\label{i4'}
2102: \Pr(\ce{(a|b)}_n=0,1,\ce{(c|d)}_n=0,1)=
2103: \Pr(\ce{(a|b)}_n=0,1)\Pr(\ce{(c|d)}_n=0,1),
2104: \end{equation}
2105:
2106:
2107: which is exactly equivalent to \eqref{i4}. The other consequences of
2108: independence are equalities
2109:
2110: \begin{align}
2111: \Pr(\ce{(a|b)}_n=1,\ce{(c|d)}=1)&=\Pr(\ce{(a|b)}_n=1)\Pr(\ce{(c|d)}=1)
2112: \tag{$\ref{i1}'$}\label{i1'}\\
2113: \Pr(\ce{(a|b)}_n=1,\ce{(c|d)}_n=0,1)&=
2114: \Pr(\ce{(a|b)}_n=1)\Pr(\ce{(c|d)}_n=0,1)
2115: \tag{$\ref{i2}'$}\label{i2'}\\
2116: \Pr(\ce{(a|b)}_n=0,1,\ce{(c|d)}_n=1)&=
2117: \Pr(\ce{(a|b)}_n=0,1)\Pr(\ce{(c|d)}_n=1),
2118: \tag{$\ref{i3}'$}\label{i3'}
2119: \end{align}
2120:
2121: which, divided by \eqref{i4'}, yield \eqref{i1}, \eqref{i2} and
2122: \eqref{i3}, respectively. Note that if both sides of \eqref{i4'} are $0,$ then
2123: all the resulting equalities involve an undefined term on both sides, and
2124: hence hold, according to our convention.
2125: \bigskip
2126:
2127: $\Uparrow$ Let \eqref{i4'} (i.e., \eqref{i4}) hold. If its both sides are $0,$
2128: the random variables $\ce{(a|b)}_n$ and $\ce{(c|d)}_n$ are independent,
2129: because one of them is constant. So let us assume \eqref{i4'} holds and its
2130: both sides are nonzero. In particular, each of the \eqref{i1}, \eqref{i2} and
2131: \eqref{i3} is defined on both sides, because the denominators are everywhere
2132: nonzero. Multiplying these equalities by \eqref{i4'}, we get \eqref{i1'},
2133: \eqref{i2'} and \eqref{i3'}, respectively. It is now a matter of routine to
2134: prove that the independence of $\ce{(a|b)}_n$ and $\ce{(c|d)}_n$ follows from
2135: \eqref{i1'},
2136: \eqref{i2'} and \eqref{i3'} and \eqref{i4'}.
2137: \end{proof}
2138:
2139: The lemma allows us to characterize independence for present tense
2140: conditionals.
2141:
2142: \begin{theorem}\label{indep_present}
2143: For present tense $(a|b)$ and $(c|d)$ the following are
2144: equivalent:
2145: \begin{itemize}
2146: \item Stochastic processes $\ce{(a|b)}$ and $\ce{(c|d)}$ are independent.
2147: \item The equalities \eqref{i1}--\eqref{i4} hold with $\pr_n$ replaced by
2148: $\Pr$ in each term, where we again assume an equation to hold in case when
2149: both sides are undefined.
2150: \end{itemize}
2151: \end{theorem}
2152:
2153:
2154: \begin{proof} For present tense $(a|b)$ the probability $\pr_n((a|b))$ is
2155: independent of $n,$ and is (of course) equal to $\Pr((a|b)).$ Now
2156: Lemmas \ref{present-tense} and \ref{indep} give us the desired
2157: equivalence.
2158: \end{proof}
2159:
2160:
2161: The full characterization of independence for general temporal
2162: conditionals is not known at the moment. Most likely, if it at all
2163: exists, it must be nonuniform, in the sense that the number of
2164: equalities between probabilities depends in principle on the actual
2165: $(\varphi|\psi)$ and $(\zeta|\xi).$
2166:
2167: However, there is a quite general sufficient condition for
2168: independence, which can be (nonuniformly) characterized by equalities
2169: of asymptotic probability.
2170:
2171:
2172: Call two conditionals $a,b$ {\em strongly independent\/} iff there
2173: exist stochastically independent Markov chains $\X$ and $\Y$ and
2174: projections $h,g$ such that $\ce{a}=h(\X)$ and $\ce{b}=g(\Y).$
2175:
2176: \begin{theorem}\label{indep_strong} Strong independence of conditional
2177: events from $\TT$ can be equivalently characterized by equations of asymptotic probability.
2178: \end{theorem}
2179:
2180: We begin with
2181:
2182: \begin{lemma}\label{MN}
2183: Let Markov chains $\X,\Y$ have $n$ and $m$ states, respectively. If
2184: $\X$ and $\Y$ are independent until time $mn+1,$ they are fully
2185: independent, i.e., if
2186:
2187: \begin{multline}\label{mn}
2188: \Pr\left(\hspace{-6pt}
2189: \begin{array}{ccc}
2190: X_1=w_1,&\dots,&X_t=w_t\\
2191: Y_1=v_1,&\dots,&Y_t=v_t
2192: \end{array}
2193: \hspace{-6pt}\right)=\\
2194: \Pr(X_1=w_1,\dots,X_t=w_t)\Pr(Y_1=v_1,\dots,Y_t=v_t)
2195: \end{multline}
2196:
2197: holds for all $t\leq mn+1$ and all sequences $w_1,\dots,w_t,$
2198: $v_1,\dots,v_t$ of states of $\X$ and $\Y,$ respectively, then $\X$
2199: and $\Y$ are independent and \eqref{mn} holds indeed for all $t.$
2200: \end{lemma}
2201: \begin{proof}
2202: First of all, observe that $(\X,\Y)=(X_1,Y_1),(X_2,Y_2),\dots$ is a
2203: Markov chain, as well.
2204:
2205: Suppose that \eqref{mn} fails and that the least $t$ for which it
2206: fails is $t>mn+1$ (because for $t\leq mn+1$ \eqref{mn} holds by
2207: assumption).
2208:
2209: The in-equality
2210:
2211: \begin{multline}\label{in-eq}
2212: \Pr\left(\hspace{-6pt}\begin{array}{ccc}
2213: X_1=w_1,&\dots,&X_t=w_t\\
2214: Y_1=v_1,&\dots,&Y_t=v_t
2215: \end{array}\hspace{-6pt}\right)\neq\\
2216: \Pr(X_1=w_1,\dots,X_t=w_t)\Pr(Y_1=v_1,\dots,Y_t=v_t)
2217: \end{multline}
2218:
2219:
2220: is by Markov property \eqref{MP} for $\X,$ $\Y$ and $(\X,\Y)$
2221: equivalent to
2222:
2223: \begin{multline*}
2224: \Pr\left(\hspace{-6pt}
2225: \begin{array}{ccc}
2226: X_1=w_1,&\dots,&X_{t-1}=w_{t-1}\\
2227: Y_1=v_1,&\dots,&Y_{t-1}=v_{t-1}
2228: \end{array}
2229: \hspace{-6pt}\right)
2230: \Pr\left(\left.\hspace{-6pt}
2231: \begin{array}{c}
2232: X_{t}=w_{t}\\
2233: Y_{t}=v_{t}
2234: \end{array}
2235: \right|
2236: \begin{array}{c}
2237: X_{t-1}=w_{t-1}\\
2238: Y_{t-1}=v_{t-1}
2239: \end{array}\hspace{-6pt}
2240: \right)
2241: \neq\\
2242: \Pr(X_1=w_1,\dots,X_{t-1}=w_{t-1})\Pr(X_t=w_t|X_{t-1}=w_{t-1})\times\\
2243: \Pr(Y_1=v_1,\dots,Y_{t-1}=v_{t-1})\Pr(Y_t=v_t|Y_{t-1}=v_{t-1}),
2244: \end{multline*}
2245:
2246: which in turn is equivalent to
2247:
2248: \begin{multline}\label{sprzecznosc}
2249: \Pr\left.\left(\hspace{-6pt}\begin{array}{c}
2250: X_{t}=w_{t}\\
2251: Y_{t}=v_{t}\end{array}\right|
2252: \begin{array}{c}
2253: X_{t-1}=w_{t-1}\\
2254: Y_{t-1}=v_{t-1}
2255: \end{array}\hspace{-6pt}\right)
2256: \neq\\
2257: \Pr(X_t=w_t|X_{t-1}=w_{t-1})
2258: \Pr(Y_t=v_t|Y_{t-1}=v_{t-1}),
2259: \end{multline}
2260:
2261: because $t$ is the least one for which in-equality holds, and so the
2262: non-conditional probabilities in the previous in-equality cancel out.
2263:
2264: Moreover, the canceling terms must be nonzero for the in-equality to
2265: hold, which means $(w_{t-1},v_{t-1})$ is reachable with positive probability
2266: from the initial state in $(\X,\Y).$ But therefore it must be
2267: reachable with positive probability in at most $mn$ steps, because
2268: there are exactly so many states in $(\X,\Y).$ So let $(x_i,y_i),\
2269: i=1,\dots, s\leq mn$ be a sequence of states of $(\X,\Y)$ leading to
2270: $(x_s,y_s)=(w_{t-1},v_{t-1})$ with positive probability. By assumption
2271:
2272: \begin{multline}
2273: \Pr\left(\hspace{-6pt}\begin{array}{ccc}
2274: X_1=x_1,&\dots,&X_s=x_s\\
2275: Y_1=y_1,&\dots,&Y_s=v_s
2276: \end{array}\hspace{-6pt}\right)=\\
2277: \Pr(X_1=x_1,\dots,X_s=x_s)\Pr(Y_1=y_1,\dots,Y_s=v_s),
2278: \end{multline}
2279:
2280: because $s\leq mn.$ If we now multiply the above by
2281: \eqref{sprzecznosc}, we get, by a calculation reverse to what we have
2282: done above, an instance of \eqref{in-eq} with $t\leq mn+1,$ a
2283: contradiction.\end{proof}
2284:
2285: \begin{lemma}
2286: For given Markov chains $\X$ and $\Y$ and for a fixed time $t,$ fixed
2287: sequences $w_1,\dots,w_t$ and $v_1,\dots,v_t$ of states of $\X$ and
2288: $\Y,$ respectively, the formula \eqref{mn} can be equivalently
2289: characterized by equalities among asymptotic probabilities of certain
2290: conditionals, derived from $\X$ and $\Y.$
2291: \end{lemma}
2292: \begin{proof} Let $\A$ and $\B$ be the deterministic finite automata,
2293: underlying $\X$ and $\Y.$ For a state $w$ of $\A$ let $\A_w$ be the Moore
2294: machine resulting from $\A$ by labeling the state $w$ with 1 and all the
2295: remaining states with 0. Since all $\A_w$'s are $\dwa$-valued, there exist
2296: $\TL$ formulas $\alpha_w,$ which are true precisely when the last symbol of
2297: the output of $\A_w$ is 1. Similarly we define $\B_v$ and $\beta_v.$
2298:
2299: Now \eqref{mn} is equivalent to
2300:
2301: \begin{multline*}
2302: \Pr(\PREV^{t}\TRUE\land\lnot\PREV^{t+1}\TRUE\land
2303: \Land_{i=1}^{t}(\PREV^{t-i}(\alpha_{w_{i}}\land\beta_{v_{i}})))=
2304: \\
2305: \Pr(\PREV^{t}\TRUE\land\lnot\PREV^{t+1}\TRUE\land
2306: \Land_{i=1}^{t}(\PREV^{t-i}(\alpha_{w_{i}})))\times\\
2307: \Pr(\PREV^{t}\TRUE\land\lnot\PREV^{t+1}\TRUE\land
2308: \Land_{i=1}^{t}(\PREV^{t-i}(\beta_{v_{i}}))).
2309: \end{multline*}
2310:
2311: Each of the $\TL$ formulas asserts that it has been once that there
2312: was something $t-1$ steps ago, but there was nothing $t$ steps ago (so
2313: we have been at time $t$ precisely), and we were in the prescribed
2314: states of the Markov chain in question at times: $t$, one step before
2315: that, \ldots, $t-1$ steps before that.
2316:
2317: \end{proof}
2318:
2319: What remains to be seen is that we can indeed choose some canonical
2320: Markov chains $\X$ and $\Y$ to represent $a$ and $b,$ which are
2321: independent whenever $a$ and $b$ are strongly independent.
2322:
2323: Let us recall, that any conditional event in our model is a projection
2324: of a Markov chain, derived from a Moore machine for the underlying
2325: conditional object. Since for every Moore machine there exists the
2326: minimal Moore machine computing the same function, in presence of
2327: probabilities, we thus always have the minimal Markov chain underlying
2328: any given conditional event.
2329:
2330: \begin{lemma}\label{minimal}
2331: Let $a$ and $b$ be strongly independent. Then the minimal Markov chains
2332: for $a$ and $b$ are independent.
2333: \end{lemma}
2334: \begin{proof} Let $\X$ and $\Y$ be two independent Markov chains, underlying
2335: $a$ and $b.$ Applying the quotient construction to $\X$ and $\Y$ we
2336: pass to the minimal Markov chains underlying $a$ and $b.$ The quotient
2337: construction is deterministic, and therefore it does not break
2338: independence (exactly like strong independence implies independence).
2339: It follows that the minimal Markov chains are independent, too.
2340: \end{proof}
2341:
2342: \begin{proof}[Proof of Theorem \ref{indep_strong}]
2343:
2344: The conditional events $a$ and $b$ are strongly independent iff the
2345: minimal Markov chains underlying them are independent, by Lemma
2346: \ref{minimal}. The latter can be expressed equivalently by
2347: $(mn+1)^{mn}$ conditions of the form \eqref{mn} for minimal chains of
2348: $m$ and $n$ states, respectively, by Lemma \ref{MN}. Each of these
2349: conditions in turn can be expressed equivalently by a single equality
2350: of asymptotic probabilities of certain conditional objects. This
2351: means that the strong independence of $a$ and $b$ can be equivalently
2352: characterized by a set of equalities among asymptotic probabilities of
2353: conditionals, which can be syntactically determined from $a$ and $b$
2354: and do not depend on the probability space structure.
2355: \end{proof}
2356:
2357: Of course, for conditionals which are themselves Markov chains for any
2358: probability assignment, strong independence is the same as independence.
2359: Therefore we have
2360:
2361: \begin{corollary} For conditionals which are themselves Markov chains
2362: for any probability assignment, independence can be characterized by
2363: equalities of asymptotic probabilities.
2364: \qed\end{corollary}
2365:
2366: The conditionals to which this applies can be recognized by the
2367: property that their minimal Moore machine has at most one state
2368: labeled by each element of $\trzy$ (and thus at most three states
2369: altogether). Present tense conditionals are of this kind, and thus we
2370: have an alternative proof of Theorem \ref{indep_present}, which much
2371: less elegant set of equalities, however. But present tense
2372: conditionals do not exhaust all conditionals, which are Markov
2373: chains. An example is the conditional $(a|\PBOX ((\PREV a\to
2374: a^\C)\land(\PREV a^\C\to a)\land(\lnot\PREV \TRUE\to a ))),$ analyzed
2375: in \cite{TT}. Its minimal Moore machine is depicted below.
2376:
2377:
2378: \begin{figure}[h!]
2379:
2380: \[\UseTips
2381: \xymatrix @C=20mm
2382: {
2383: &*+++[o][F]{1}
2384: \ar@/^/[dr]^{\textstyle{a}}
2385: \ar@<-.4em>[dd]_{\textstyle{a^\C}}
2386: \\
2387: &&*+++[o][F]{\bot}
2388: \ar@(ur,dr)[]
2389: \\
2390: \ar[r]
2391: &*+++[o][F]{0}
2392: \ar@/_/[ur]_{\textstyle{a^\C}}
2393: \ar@<-.4em>[uu]_{\textstyle{a}}
2394: }
2395: \]
2396:
2397: \caption[Non-simple conditional with 3-states]{Moore machine of
2398: $(a|\PBOX ((\PREV a\to a^\C)\land(\PREV a^\C\to a)\land(\lnot\PREV
2399: \TRUE\to a ))).$}\label{f_indep}
2400: \end{figure}
2401:
2402:
2403: Therefore Theorem \ref{indep_strong} is indeed stronger than Theorem
2404: \ref{indep_present}.
2405:
2406: Finally, we consider the question of $\PS$ \cea, for which one might
2407: want a characterization of independence in terms of asymptotic
2408: probability. Here we give a negative answer.
2409:
2410: \begin{theorem}\label{(a|0)} There exist two conditional expressions
2411: $e_1$ and $e_2$ and a probability space such that the embeddings
2412: $\ttt{e_1}$ and $\ttt{e_2}$ are independent, while their sparse reverse
2413: counterparts $\ttrs{e}$ and $\ttrs{e_2}$ are not independent.
2414: \end{theorem}
2415: \begin{proof} Take $e_1=e_2=(0|a)$ and any probability space with
2416: $0<\Pr(a)<1.$ Then $\ce{\ttt{e_1}}$ and $\ce{\ttt{e_2}}$ are constant
2417: processes, equal to $0,$ so they are (trivially) independent. However,
2418: already
2419:
2420: \begin{align*}\Pr(\begin{array}{c}\ce{\ttrs{e_1}})(w)=0\bot
2421: \\\ce{\ttrs{e_2}})(w)=0\bot\end{array})&=
2422: \Pr(\ce{\ttrs{e_1}})(w)=0\bot)\\
2423: &>(\Pr((\ce{\ttrs{e_1}})(w)=0\bot))^2\\
2424: &=\Pr(\ce{\ttrs{e_1}})(w)=0\bot)\cdot
2425: \Pr(\ce{\ttrs{e_2}})(w)=0\bot),
2426: \end{align*}
2427:
2428: where the inequality holds because
2429: $\Pr(\ce{\ttrs{e_1}})(w)=0\bot)=\Pr(a)(1-\Pr(a))\neq 0,1.$\end{proof}
2430:
2431:
2432: \begin{corollary} \label{PS-indep} There is no characterization of
2433: independence in $\PS$ \cea\ in terms of equalities of asymptotic
2434: probability.
2435: \end{corollary}
2436: \begin{proof} Because both $\ttt{\cdot}$ and $\ttrs{\cdot}$ preserve
2437: all asymptotic probabilities of conditionals, both of them satisfy
2438: precisely the same equalities of asymptotic probabilities. So if there
2439: were a characterization of independence in terms of equalities of such
2440: probabilities, the interpretations of the two conditionals $(a|0)$ and
2441: $(a|0)$ above would have to be either independent in both cases, or
2442: dependent in both cases, while they are not, a contradiction.
2443: \end{proof}
2444:
2445:
2446:
2447: The consequence is that in the \cea{} formalism is not expressive
2448: enough to define independence of conditionals by means of equalities
2449: of asymptotic probabilities. Note however, that such a representation
2450: is certainly possible by means of equalities of probabilities and
2451: equalities of the algebraic structure. Indeed, $\PS$ \cea{} is boolean
2452: algebra with respect to its connectives $\land,\lor,\llnot$ (as it is
2453: easily visible from its syntactic representation within $\TT$), and
2454: the equalities it satisfies enforce, that $\Pr_o$ is an ordinary
2455: probability measure. Therefore independence is equivalent to the
2456: standard equality $\Pr_o(\ps{e_1\land
2457: e_2})=\Pr_o(\ps{e_1})\Pr_o(\ps{e_2}).$ What we have constructed are
2458: two non-boolean subsystems of $\TT$, in which all the (asymptotic)
2459: probability assignments agree with those of $\Pr_o$, and yet no set of
2460: equalities of probabilities can characterize the true probabilistic
2461: independence in both of them simultaneously.
2462:
2463: \subsection{Algorithms}\label{algorytmy}
2464:
2465: \paragraph{Polynomial algorithm for PS \cea.} Let us see that our
2466: approach provides a nontrivial improvements to the algorithmic status of
2467: existing \cea's. We will demonstrate this by calculating the
2468: probabilities of conditional expressions, according to $\PS$ \cea, in
2469: time polynomial in their size and exponential in the number of
2470: variables. (Note that the number of arguments for computation of the
2471: probability of an $n$-ary conditional is $2^n,$ so the above indeed
2472: means computation polynomial in the size of the input.) In \cite{g94} it
2473: is stated that the computation of the $\PS$-probability of a conjunction
2474: of $n$ conditionals $(a_i| b_i),$ according to the method used in that
2475: paper, requires adding $\sum_{m=1}^n m!\cdot S_0(m,n)\cdot (2^{m+1}-2)$
2476: terms, each being a nonconditional probability of a conjunction of
2477: certain events $a_i$ and $b_i.$ The number of summands, where $S_0(m,n)$
2478: are Stirling's number of the second kind, is of order $2^{n\log n}.$ It
2479: is substantially more than about $c2^{n}$ one obtains for the present
2480: tense \cea's $\SAC$ and $\GNW,$ and has been stressed in
2481: \cite[p.~499]{kniga} and in \cite{recenzja}, since it strongly affects
2482: the usefulness of $\PS$ as a tool for applications. Using our approach we
2483: have instantly an algorithm to calculate the same probability in
2484: $2^{O(n)}$ steps. As a matter of fact, this applies to {\em any\/}
2485: conditional expression with $n$ arguments $(a_i| b_i),\ i=1\dots,n,$ as
2486: long as its length does not exceed $2^{O(n)}.$ All the complexity bounds
2487: given here assume unit cost of basic arithmetical operations: addition,
2488: multiplication, subtraction and division.
2489:
2490: \begin{theorem} There is an algorithm, computing the $\PS$\/ probability
2491: of an $n$-ary conditional expression of length $m$ in time polynomial in
2492: $\max(m,2^n).$
2493: \end{theorem}
2494: \begin{proof} The first step of the algorithm on input expression $e$ is
2495: to construct the minimal Moore machine, computing $\ttt{e}.$
2496:
2497: \begin{lemma} The minimal Moore machine of $\ttt{e}$ for $n$-ary
2498: conditional expression $e$ has at most $3^n$ states.
2499: \end{lemma}
2500: \begin{proof} The minimal Moore of $\first(a|b)$ has 3 states (see
2501: \eqref{first} and Figure \ref{f4}).
2502:
2503: By the definition of the $\ttt{\cdot}$ embedding (Section \ref{FE}), a
2504: Moore machine $\A=(Q,\Omega,\delta,h,q_0)$ of $\ttt{e}$ can be
2505: constructed as follows:
2506:
2507: The set $Q$ of states of $\A$ is the product $Q_1\times\dots\times Q_n$
2508: of state sets of Moore machines $\A_i=(Q_i,\Omega,\delta_i,h_i,q_{0i})$
2509: of all expressions $\first(a_i|b_i)$ occurring in $\ttt{e}.$ The transition
2510: function of $\A$ is defined coordinate-wise, i.e.,
2511:
2512: \[\delta(\langle q_1,\dots,q_n\rangle,\omega)=\langle
2513: \delta_1(q_1,\omega),\dots,\delta_n(q_n,\omega)\rangle,\]
2514:
2515: the initial state is $q_0=
2516: \langle q_{01},\dots,q_{0n}\rangle,$ and, crucially,
2517:
2518: \[h(\langle
2519: q_{1},\dots,q_{n}\rangle)=\hat{c}(h_1(q_1),\dots,h_n(q_n)),\]
2520:
2521: where $\hat{c}(h_1(q_1),\dots,h_n(q_n))$ is the classical logic
2522: evaluation of the expression $c$ on arguments
2523: $h_1(q_1),\dots,h_n(q_n)\in\dwa.$
2524:
2525: This product construction is well-known for automata theory, and it is
2526: immediate that it does the work.
2527: \end{proof}
2528:
2529: So it is quite easy to construct, given $e\in\LL,$ the Moore machine of
2530: $\ttt{e}.$ Now we have to turn this Moore machine into a Markov chain.
2531: Assuming that all the probabilities of atomic events from $\Omega$ are
2532: given, we simply replace multiple transitions between the same states
2533: represented by the sum of their probabilities---a single number.
2534:
2535: Furthermore, the Markov chain we obtain is absorbing, i.e., it has
2536: one-element ergodic classes. It can be proven by a straightforward
2537: induction on $n$ --- the number of three element Moore machines we
2538: product. It follows \cite[Chapter III]{KS} that we can use the following
2539: method to compute the limiting probability that the chain finally
2540: arrives at a state labeled by $1$.
2541:
2542: Clearly, in this situation we can collapse all absorbing states labeled
2543: $1$ into a single such state.
2544:
2545: Denote by $P$ the matrix $(p(i,j))$ of transition probabilities, by $Q$
2546: the submatrix of rows and columns corresponding to transient states, and
2547: by $R$ the submatrix of rows corresponding to transient states and
2548: columns corresponding to absorbing states. Let $Id$ be a diagonal matrix
2549: with $1$'s on the diagonal and $0$'s elsewhere. Let $B=(Id-Q)^{-1}R.$
2550: Then the probability we are looking for is the entry in $B$ in the row
2551: corresponding to the initial state in in the column corresponding to the
2552: (only) absorbing state labeled by $1$ in the Markov chain. Since all
2553: the calculations on matrices necessary to compute $B$ are doable in time
2554: polynomial in the size of the matrices, the total computation time is
2555: $(2^{n})^{O(1)}=2^{O(n)},$ as desired.
2556: \end{proof}
2557:
2558:
2559: \section{Summary}
2560:
2561: We have discussed the temporal calculus of conditional objects and
2562: conditional events $\TT$ as a formalism alternative to conditional event
2563: algebras.
2564:
2565: We have shown that all the major conditional event algebras, including
2566: those of Schay-Adams-Calabrese, Goodman-Nguyen-Walker and the product
2567: space \cea, embed isomorphically in $\TT.$
2568:
2569: Moreover, $\TT$ is superior to those formalisms in several ways:
2570:
2571: \begin{itemize}
2572: \item It provides natural, probabilistic semantics of conditionals,
2573: allowing one to construct experiments to evaluate all their interesting
2574: probabilistic parameters, unlike \cea's, which generally are not
2575: probability spaces, and which do not require certain probabilistic
2576: parameters to be defined at all.
2577: \item The construction of $\TT$ is functorial, in the sense, that the
2578: underlying probabilistic space of nonconditional events determines the
2579: space of temporal conditional events uniquely, while \cea's generally
2580: are not unique.
2581: \item The formalism of $\TT$ allows one to define and analyze
2582: independence of conditional events, which is difficult or impossible in
2583: \cea's.
2584: \item $\TT$ offers better algorithms for calculation of probabilities,
2585: than those known previously for \cea's.
2586: \end{itemize}
2587:
2588: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2589: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2590:
2591: \bibliographystyle{alpha} \bibliography{bib}
2592:
2593: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2595:
2596: \end{document}
2597: