cs0110003/tt.tex
1: \NeedsTeXFormat{LaTeX2e}[1994/06/01]
2: \documentclass[11pt,a4paper]{article} 
3: 
4: \usepackage{latexsym}
5: \usepackage{amssymb} 
6: \usepackage{amsmath} 
7: \usepackage{amsthm}
8: \usepackage{eucal} 
9: %\usepackage{maplesty}
10: \usepackage{bezier}
11: \usepackage{graphics}
12: \usepackage[matrix,curve,arrow,tips,frame]{xy}
13: \usepackage{color}
14: 
15: \theoremstyle{plain} \newtheorem{theorem}{Theorem}
16: \newtheorem{lemma}[theorem]{Lemma}
17: \newtheorem{sublemma}[theorem]{Sublemma}
18: \newtheorem{proposition}[theorem]{Proposition}
19: \newtheorem{corollary}[theorem]{Corollary} \newtheorem{cl}{Claim}
20: \newtheorem{conj}{Conjecture} \newtheorem*{observation}{Observation}
21: 
22: \theoremstyle{definition} \newtheorem{definition}[theorem]{Definition}
23: \newtheorem{algorithm}{Algorithm}
24: \renewcommand{\thealgorithm}{\Alph{algorithm}}
25: 
26: \theoremstyle{remark} \newtheorem*{claim}{Claim}
27: \newtheorem*{example}{Example} 
28: \newtheorem*{acknowledgement}{Acknowledgement}
29: 
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 &#2 &#3 &#4\\ 1 &#5 &#6 &#7\\ \bot &#8 &#9
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 &#2 &#3 &#4\\ \hline 1 &#5 &#6 &#7\\ \hline \bot &#8 &#9
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: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161: \title{The Temporal Calculus\\ of Conditional Objects and Conditional
162: Events}
163: 
164: \author{
165: Jerzy Tyszkiewicz$^{1,2}$\\
166: Arthur Ramer$^2$\\
167: Achim Hoffmann$^2$}
168: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
170: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
171: \CompileMatrices
172: \pagestyle{myheadings}
173: \markright{{\tt \jobname .tex}, version of \today}
174: \begin{document} 
175: \begin{titlepage}
176: \maketitle 
177: \begin{center}$^1$ Institute of Informatics,\\ University of
178: Warsaw,\\ Banacha 2,\\ 02-097 Warszawa,\\ Poland.\\ E-mail
179: {\tt jty@mimuw.edu.pl}.\\ Supported by the Polish Research Council KBN
180: grant 8 T11C 027 16.\\[5pt]  $^2$ School CSE,\\ UNSW,\\ 2052 Sydney,\\
181: Australia.\\ E-mail {\tt \{jty|ramer|achim\}@cse.unsw.edu.au}.\\ Supported by
182: the Australian Research Council ARC grant A 49800112 (1998--2000).
183: \end{center}
184: \thispagestyle{empty}
185: \end{titlepage}
186: \begin{titlepage}
187: \vspace{5cm}
188: \begin{abstract} 
189:     We consider the problem of defining conditional objects $(a|b),$
190: which would allow one to regard the conditional probability $\Pr(a|b)$
191: as a probability of a well-defined event rather than as a shorthand
192: for $\Pr(ab)/\Pr(b).$ The next issue is to define boolean combinations
193: of conditional objects, and possibly also the operator of further
194: conditioning.  These questions have been investigated at least since
195: the times of George Boole, leading to a number of formalisms proposed
196: for conditional objects, mostly of syntactical, proof-theoretic vein.
197: 
198:     We propose a unifying, semantical approach, in which conditional
199: events are (projections of) Markov chains, definable in the
200: three-valued extension $\TT$ of the past tense fragment of
201: propositional linear time logic ($\TL$), or, equivalently, by
202: three-valued counter-free Moore machines. Thus our conditional objects
203: are indeed stochastic processes, one of the central notions of modern
204: probability theory. 
205: 
206: Our model precisely fulfills early ideas of de Finetti \cite{d72},
207: and, moreover, as we show in a separate paper \cite{cea}, all the
208: previously proposed algebras of conditional events can be
209: isomorphically embedded in our model.
210: \end{abstract}
211: \end{titlepage}
212: 
213: \begin{titlepage}
214: \tableofcontents 
215: \end{titlepage}
216: 
217: \section{Preliminaries and statement of the problem}
218: 
219: \subsection{The problem of conditional objects}
220: 
221: Probabilistic reasoning \cite{p88} is the basis of Bayesian methods of
222: expert system inferences, of knowledge discovery in databases, and in
223: several other domains of computer, information, and decision
224: sciences.  The model of conditioning and conditional objects we discuss
225: serves equally to reason about probabilities over a finite domain $X$,
226: or probabilistic propositional logic with a finite set of atomic
227: formulae.
228: 
229: 
230: Computing of conditional probabilities of the form
231: $\Pr(X|Y_1,\dots,Y_n)$ and, by extension of conditional beliefs, is
232: well understood.  Attempts of defining first the {\em conditional
233: objects\/} of the basic form $X|Y$, and then defining $\Pr(X|Y)$ as
234: $\Pr((X|Y))$ were proposed, without much success, by some of the
235: founders of probability \cite{b57,d72}.  They were taken up
236: systematically only about 1980.  The development was slow, both
237: because of logical difficulties \cite{l76,cccp1,cccp2}, and even more
238: because the computational model is difficult to construct.  (While
239: $a|b$ appears to stand for a sentence `if $b$ then $a$', there is no
240: obvious calculation for $\Pr(a|(b|c))$, nor intuitive meaning for
241: $a|(b|c),$ $(a|b)\land(c|d),$ and the like.)
242: 
243: The idea of defining conditional objects was entertained by some
244: founders of modern probability \cite{b57,d72}, but generally abandoned
245: since introduction of the measure-theoretic model.  It was revived
246: mostly by philosophers in 1970's \cite{a86,v77} with a view towards
247: artificial intelligence reasoning.  Formal computational models came
248: in the late 1980's and early 1990's \cite{c87,gnw91,g94}. Only a few
249: of them have been used for few actual calculations of conditionals and
250: their probabilities whose values are open to questions \cite{c94,g94}.
251: 
252: In this paper we want to give a rigorous (and yet quite natural and
253: intuitive) probabilistic and semantical construction of conditionals,
254: based on ideas proposed by de Finetti over a quarter a century ago
255: \cite{d72}. It appears that this single formalism contains fragments
256: precisely corresponding to all the previously considered algebras of
257: conditional events \cite{cea}.  Seen as a whole, it can be therefore
258: considered as their common generalisation and perhaps {\em the\/}
259: calculus of conditionals. 
260: 
261: Our system consists of three layers: the logical part is a three
262: valued extension of the past tense fragment of propositional linear
263: time logic, the computation model are three-valued Moore machines (an
264: extension of deterministic finite automata), and the probabilistic
265: semantics is provided by three-valued stochastic processes, which
266: appear to be projections of Markov chains.
267: 
268: 
269: \subsection{The main idea}
270: 
271: \paragraph{The main idea.} The main idea of our approach can be seen
272: as an attempt to provide a precise mathematical implementation of the
273: following idea of de Finetti \cite[Sect.\ 5.12]{d72}:
274: 
275: \begin{quote}
276: \begin{sl}
277: ``In the asymptotic approach, the definition of conditional
278: probability appears quite naturally; it suffices to repeat the
279: definition of probability \textup{(}as the limiting
280: frequency\textup{)}, taking into consideration only the trials in
281: which the conditioning event \textup{(}hypothesis\textup{)} is
282: satisfied. Thus, $P(E|H)$ is simply the limit of the ratio between the
283: frequency of $EH$ and the frequency of $H.$ If the limiting frequency
284: of $H$ exists and is different from zero, the definition is
285: mathematically equivalent to the compound probability theorem
286: $P(E|H)=P(EH)/P(H).$ But even if the frequency of $H$ does not tend to
287: a limit, or the limit is zero, $P(E|H)$ can nonetheless exist
288: \textup{(}trivial example: $P(H|H)$ is always equal to \/
289: $1$\textup{)}.''
290: \end{sl}
291: \end{quote}
292: 
293: We believe that our attempt is successful: our system will have all
294: the properties predicted by de Finetti, and, moreover, as we show in a
295: separate paper \cite{cea}, subsumes all the previously existing
296: formalisms developed to deal with conditionals, and, finally, appears
297: to be able to handle some well-known paradoxes of probability in an
298: intuitive and yet precise manner.
299: 
300: \paragraph{Three truth values.} 
301: 
302: To be able to take into account only the trials in which the
303: hypothesis is satisfied, one has to introduce a third logical
304: value. Informally, if one considers two players\footnote{This sounds
305: definitely better than gamblers ;-).}: one betting $(a|b)$ will hold,
306: and the other it will not, if in a random experiment (dice toss, coin
307: flip) $b$ doesn't hold, the game is drawn.  The previous works
308: considered it to be an evidence that the definition of conditionals
309: must be necessarily based on many valued logics, the typical choices
310: being three valued.
311: 
312: Note however, that assigning probability to a three-valued $c$ is
313: something like squeezing it to become two-valued. For one then assumes
314: it to be true $\Pr(c)$ of time and false $1-\Pr(c)$ of time, and the
315: time when $c$ has the third value, typically described as {\em
316: undefined}, is lost. So, unlike most of our predecessors, we attempt
317: to preserve the three-valuedness of conditionals as a principle, and
318: define their probability only on the top of that.
319: 
320: 
321: \paragraph{Bet repetitions.}
322: Now, we should allow the players to repeat their bets. Here, unlike
323: most of the previous works, if the players repeat the game, we allow
324: them to bet on properties of the {\em whole\/} sequence of outcomes,
325: not just the last one.
326: 
327: This is not uncommon in many random experiments, that the history of
328: the bets influences the present bet somehow.
329: 
330: We present three natural examples, which are natural and have a simple
331: description.  
332: 
333: The first possibility is that after each bet we we start over---after
334: the result of the experiment is settled, the (temporal) history is
335: started anew, the next experiment not taking the old results into
336: account.
337: 
338: The second is just the opposite---always the entire history, including
339: earlier experiments, is taken into account.
340: 
341: The third is that no repetition is allowed: after the first experiment
342: is settled, its outcome is deemed to persist forever, and future
343: trials are effectively null.  (Regardless of each subsequent element
344: drawn the result is always defined and remains the same.)
345: 
346: Roughly speaking, the first choice is adopted in bridge, the second in
347: blackjack and the third in Russian roulette.
348: 
349: 
350: This suggests that a conditional isn't merely an experiment with three
351: possible outcomes. It is indeed a {\em sequence} of experiments, and
352: the third logical value, often described as {\em unknown}, is often
353: {\em not y\underline{et} known}. It is clearly a temporal concept, and
354: thus we are going to consider conditionals as {\em temporal objects.}
355: This temporal aspect is clearly of {\em past tense\/} type --- the
356: result of a bet must depend on the history (including present) of the
357: sequence of outcomes, only.
358: 
359: It is worth noting that there are other approaches which consider
360: implicitly bet repetition in the modelling of conditionals. These
361: include \cite{v77,mcgee,g94,ramer}.
362: 
363: \paragraph{Summary.} 
364: What we undertake is thus the development of a calculus of conditional
365: objects identified with temporal rules, which, given a sequence of random
366: elements from the underlying domain, decide after each of the drawn
367: elements if the the conditional becomes defined, and if so, whether it
368: is true or false.
369: 
370: We stipulate that, for any reasonable calculus of conditionals, {\em
371: forming boolean combinations of conditionals, as well as iterated
372: conditionals, amounts to manipulating on these rules.}
373: 
374: This claim is indeed well motivated: if we fail to associate such rule
375: to a complex conditional object, we do not have any means to say, in a
376: real-life situations, who wins the bet on this conditional and when.
377: So to say, such a conditional would be nonprobabilistic, because one
378: couldn't bet on it!
379: 
380: \paragraph{Novelty of our approach.}  We would like to stress that
381: virtually none of the results we prove below is entirely new. Most of
382: them are simple extensions or reformulations of already known
383: theorems, as the reader can verify in Section \ref{related}. The
384: novelty of our approach lies almost entirely in the way we assemble
385: the results to create mathematically precise representation of an
386: otherwise quite clear and intuitive notion. And indeed, we feel very
387: reassured by the fact that we didn't have to invent any new
388: mathematics for our construction.  Similarly the proofs we give in
389: this paper are quite straightforward. This is exactly the emergence of
390: previously-unheard-of complicated algebraic structures (dubbed {\em
391: conditional event algebras\/} in \cite{kniga}), which prompted us to
392: have a closer look at conditional events and search for simpler and
393: more intuitive formalisations. Note that probabilists and logicians
394: have been doing quite well without conditional events for decades,
395: which strongly suggests they have had all the tools necessary to use
396: conditionals in an implicit way for a long time already. To the
397: contrary, in the emerging applied areas, and in particular in AI,
398: there is a strong need to have conditional events explicitly present,
399: and this is why we believe in the importance of our results.
400: 
401: 
402: \section{The tools}
403: \subsection{Pre-conditionals}
404: 
405: Let $\E=\{a,b,c,d,\dots\}$ be a finite set of basic events, and let
406: $\Sigma$ be the free Boolean algebra generated by $\E,$ and $\Omega$
407: the set of atoms of $\Sigma.$ Consequently, $\Sigma$ is isomorphic to
408: the powerset of $\Omega,$ and $\Omega$ itself is isomorphic to the
409: powerset of $\E.$ Any element of $\Sigma$ will be considered as an
410: event, and, in particular, $\E\subseteq\Sigma.$
411: 
412: The union, intersection and complementation in $\Sigma$ are denoted by $a\cup
413: b,\ a\cap b$ and $a^\C,$ respectively.  The least and greatest elements of
414: $\Sigma$ are denoted $\varnothing$ and $\Omega,$ respectively. However,
415: sometimes we use a more compact notation, replacing $\cap$ by juxtaposition.
416: When we turn to logic, it is customary to use yet another notation: $a\lor b,\
417: a\land b$ and $\lnot a,$ respectively. In this situation $\Omega$ appears as
418: $\TRUE$ and $\varnothing$ as $\FALSE,$ but $1$ and $0,$ respectively, are
419: incidentally used, as well. Generally we are quite anarchistic in our
420: notation, as long as it does not create ambiguities.
421: 
422: We introduce the set $\trzy =\{0,1,\bot\}$ of truth values,
423: interpreted as {\em true}, {\em false\/} and {\em undefined},
424: respectively.  The subset of $\trzy $ consisting of $0$ and $1$ will
425: be denoted $\dwa.$
426: 
427: 
428: It follows from the discussion above that we are going to look for
429: conditionals in the set $\PC=\trzy ^{\Omega^+}$ of three-valued
430: functions $c$ from the set $\Omega^+$ of finite nonempty sequences of
431: atomic events from $\Omega$ into $\trzy .$ We will call such functions
432: {\em pre-conditionals}, since to deserve the name of conditionals they
433: must obey some additional requirements.
434: 
435: Sometimes it is convenient to represent such objects in two other, slightly
436: different, yet equivalent forms: 
437: 
438: \begin{itemize}
439: \item The second representation are length-preserving mappings
440: $c_+:\Omega^+\to\trzy ^+$ such that $c_+(v)$ is a prefix of $c_+(vw).$
441: The set of all such mappings will be denoted $\PC_+.$
442: \item The third representation are mappings
443: $c_\infty:\Omega^\infty\to\trzy ^\infty$ such that if
444: $w,v\in\Omega^\infty$ have a common prefix of length $n,$ then
445: $c_\infty(w)$ and $c_\infty(v)$ have a common prefix of length $n,$
446: too.  The set of all such mappings will be denoted $\PC_\infty.$
447: \end{itemize}
448: 
449: On the set $\Omega^+\cup\Omega^\infty$ one has the natural partial
450: order relation of being a prefix. Suprema of sets in this partial
451: order are denoted by $\bigsqcup.$
452: 
453: In general, $c,\ c_+$ and $c_\infty$ denote always three
454: representations of the same pre-conditional, and the subscript (or its
455: lack) indicates what representation we take at the moment, and we
456: choose it according to what is most convenient. The three
457: representations $c,\ c_+$ and $c_\infty$ are related by the equalities
458: 
459: \begin{equation}\label{reprezentacje}
460: \begin{split}
461: c(\omega_1\dots \omega_n)&=\text{last-letter-of}(c_+(\omega_1\dots
462: \omega_n)),\\
463: c(\omega_1\dots \omega_n)&=\text{$n$th-letter-of}(c_\infty(\omega_1\dots
464: \omega_n\dots)),\\
465: c_+(\omega_1\dots \omega_n)&=c(\omega_1)c(\omega_1\omega_2)\dots
466: c(\omega_1\dots \omega_n),\\ 
467: c_+(\omega_1\dots \omega_n)&=\text{first-$n$-letters-of}
468: (c_\infty(\omega_1\dots\omega_n\dots)),\\ 
469: c_\infty(\omega_1\dots \omega_n\dots)&=c(\omega_1)c(\omega_1\omega_2)\dots
470: c(\omega_1\dots \omega_n)\dots,\\ 
471: c_\infty(\omega_1\dots \omega_n\dots)&=\bigsqcup\{c_+(\omega_1\dots
472: \omega_n)~/~n=1,2\dots\}.
473: \end{split}
474: \end{equation}
475: 
476: 
477: 
478: Even though we are on a rather preliminary level of our construction, we
479: can address the general question of defining connectives among
480: pre-conditionals already now. In our setting such a connective is indeed
481: a function from some power of the space of pre-conditionals into itself.
482: However, to fulfill the requirement that a connective should depend
483: solely on the outcomes of its arguments (this property is called {\em
484: extensionality\/} in the logic literature), and that it should refer to
485: the history, only, the following additional condition must be met.
486: 
487: For any connective $\alpha:\PC_+^n\to \PC_+$
488: and any $\varphi_1,\dots,\varphi_n,\varphi'_1,\dots,\varphi'_n\in
489: \PC_+,$ $v,w\in\Omega^+$ satisfying $\varphi_i(w)=\varphi'_i(v)$ for
490: $i=1,\dots,n$ holds
491: 
492: \[\alpha(\varphi_1,\dots,\varphi_n)(w)=
493: \alpha(\varphi'_1,\dots,\varphi'_n)(v).\]
494: 
495: Note that we permit strong dependence on the history: we do not
496: require the connective to depend just on the present values of its
497: arguments, we allow it to depend on their whole histories. However, if
498: a particular connective $\alpha$ meets the former, stronger
499: requirement, whose formal statement can be obtained from the above
500: condition by replacing $\PC_+$ by $\PC$ everywhere it occurs, we call
501: it a {\em present tense connective}.
502: 
503: Connectives which are not present tense will be called {\em past
504: tense}. Any $n$-ary present tense connective of pre-conditionals is
505: fully characterised by a mapping $\trzy ^n\to \trzy.$ Note that any
506: connective $\alpha,$ not necessarily present tense one, can be
507: completely specified by a mapping $\bigcup_{t>0}\underbrace{\trzy
508: ^t\times\dots\times\trzy ^t}_{\text{$n$ times}}\to \trzy .$
509: 
510: Just like their connectives, pre-conditionals can be present tense,
511: too. A pre-conditional $c:\Omega^+\to\trzy $ is called {\em present
512: tense\/} iff $c(v)=c(w)$ holds whenever
513: $\text{last-letter-of}(v)=\text{last-letter-of}(w).$ So indeed a
514: present tense pre-conditional is completely determined by a function
515: $\Omega\to\trzy .$
516: 
517: \subsection{The formalisms}
518: 
519: 
520: Our intention is to distinguish conditionals among pre-conditionals.
521: Therefore, in order to deal with them, we need a formalism aimed at
522: dealing with sequences of symbols from a finite alphabet. There are
523: many candidates of this kind, including regular expressions and their
524: subclasses, grammars of various kinds, deterministic or
525: nondeterministic automata, temporal logics, first order logic and
526: higher order logics.
527: 
528: 
529: Our choice, which will be carefully motivated later on, is to use
530: three-valued counterparts of a certain particular class of finite
531: automata and of past tense temporal logic. When the probabilities come
532: into play conditional events of a fixed probability space are
533: represented by Markov chains.
534: 
535: We introduce here briefly the main formalisms used throughout this
536: paper: temporal logic, Moore machines and Markov chains.
537: 
538: \subsection{Temporal logic}
539: 
540: Let us first define {\em temporal logic of linear discrete past time},
541: called $\TL.$ We follow the exposition in \cite{Emerson}, tailoring
542: the definitions somewhat towards our particular needs.
543: 
544: The formulas are built up from the set $\E$ (the same set of basic
545: events as before), interpreted as propositional variables here, and
546: are closed under the following formula formation rules:
547: 
548: \begin{enumerate}
549: \item Every $a\in \E$ is a formula of temporal logic.
550: \item If $\varphi,\psi\in \TL,$ then their boolean combinations
551: $\varphi\lor \psi$ $\lnot\varphi$ are in $\TL.$ The other Boolean
552: connectives: $\land,\to,\lra,\dots$ can be defined in terms of $\lnot$
553: and $\lor,$ as usual.
554: 
555: \item If $\varphi,\psi\in \TL,$ then their past tense temporal
556: combinations $\PREV\varphi$ and $\varphi\Since\psi$ are in $\TL,$
557: where $\PREV\varphi$ is spelled ``previously $\varphi.$''
558: \end{enumerate}
559: 
560: A model of temporal logic is a sequence $\M=s_0,s_1,\dots,s_n$ of states, each
561: state being a function from $\E$ (the same set of basic events as before) to
562: the boolean values $\{0,1\}.$ Note that a state can be therefore understood as
563: an atomic event from $\Omega,$ and $\M$ can be thought of as a word from
564: $\Omega^+.$ To be explicit we declare that the states of
565: $\M$ are ordered by $\leq.$ Rather than using the indices of states to denote
566: their order, we simply write $s\leq t$ to denote that a state $t$ comes later
567: than, or is equal to, a state $s;$ similarly $s+1$ denotes the successor state
568: of $s.$ We adopt the convention that, unless explicitly indicated otherwise, a
569: model is always of length $n+1,$ and thus $n$ is always the last state of a
570: model.
571: 
572: For every state $s$ of $\M$ we define inductively what it means that a
573: formula $\varphi\in\TL$ is satisfied in the state $s$ of $\M,$
574: symbolically $\M,s\models\varphi.$
575: 
576: \begin{enumerate}
577: \item $\M,s\models a$ iff $s(a)=1$
578: \item
579: \begin{align*}
580: \M,s\models\lnot\varphi&:\iff\ \M,s\not\models\varphi,\\
581: \M,s\models\varphi\lor\psi&:\iff
582: \M,s\models\varphi\ \text{or}\ \M,s\models\psi.
583: \end{align*}
584: 
585: 
586: \item
587: \begin{align*}
588: \M,s\models\PREV\varphi&:\iff s>0\ \text{and}\ \M,s-1\models\varphi;\\
589: \M,s\models\varphi\Since\psi&:\iff(\exists t\leq s)(\M,t\models\psi\
590: \text{and}\ (\forall t< w\leq s)M,w\models\varphi).
591: \end{align*}
592: \end{enumerate}
593: 
594: The syntactic abbreviations $\PBOX\varphi$ and $\PDIA\varphi$ are of
595: common use in $\TL.$ They are defined by $\PDIA\varphi\equiv\FALSE
596: \Since\varphi$ and $\PBOX\varphi\equiv\lnot\PDIA\lnot\varphi.$ The
597: first of them is spelled ``once $\varphi$'' and the latter ``always in
598: the past $\varphi$''.
599:  
600: Their semantics is then equivalent to
601: \begin{align*}
602: \M,s\models\PBOX\varphi&:\iff(\forall t\leq s)\M,t\models\varphi;\\
603: \M,s\models\PDIA\varphi&:\iff(\exists t\leq s) \M,t\models\varphi.
604: \end{align*}
605: 
606: 
607: Using the given temporal and boolean connectives, one can write down
608: quite complex formulae describing temporal properties of models
609: $\M,s.$ We will see several such examples in this paper, and even more
610: can be found in  \cite{cea}.
611: 
612: 
613: \subsection{Moore machines}\label{MM}
614: 
615: In this section we follow \cite{HU}, tailoring the definitions, again,
616: towards our needs.
617: 
618: 
619: A {\em deterministic finite automaton\/} is a five-tuple
620: $\A=(Q,\Omega,\delta,q_0,T),$ where $Q$ is its set of states, $\Omega$
621: (the same set of atomic events as before) is the input alphabet,
622: $q_0\in Q$ is the initial state and $\delta: Q \times\Omega\to Q$ is
623: the transition function.  $T\subseteq Q$ is the set of accepting
624: states.
625: 
626: We picture $\A$ as a labelled directed graph, whose vertices are
627: elements of $Q,$ a the function $\delta$ is represented by directed
628: edges labelled by elements of $\Omega$: the edge labelled by
629: $\omega\in\Omega$ from $q\in Q$ leads to $\delta(q,\omega).$ The
630: initial state is typically indicated by an unlabelled edge ``from
631: nowhere'' to this state.
632: 
633: As the letters of the input word $w\in\Omega^+$ come in one after
634: another, we walk in the graph, always choosing the edge labelled by
635: the letter we receive. What we do with the word depends on the state
636: we are in upon reaching  the end of the word. If it is in $T,$ the
637: automaton accepts the input, otherwise it rejects it.
638: 
639: Formally, to describe the computation of $\A$ we extend $\delta$ to a
640: function $\hat{\delta}: Q \times\Omega^+\to Q$ in the
641: following way: 
642: 
643: \[\hat{\delta}(q,w)=\begin{cases}
644: \delta(q,w)&\text{if $|w|=1$}\\
645: \delta(\hat{\delta}(q,v),\omega)&\text{if $w=v\omega.$}
646:                     \end{cases}
647: \]
648: 
649: $L(\A)\subseteq\Omega^+$ is the set of words accepted by $\A.$
650: 
651: A {\em Moore machine $\A$} is a six-tuple
652: $\A=(Q,\Omega,\Delta,\delta,h,q_0),$ where $(Q,\Omega,\delta,q_0)$ is
653: a deterministic finite automaton but the set of accepting states,
654: $\Delta$ is a finite output alphabet and $h$ is the output function
655: $Q\to\Delta.$ In addition to what $\A$ does as a finite automaton, at
656: each step it reports to the outside world the value $h(q)$ of the
657: state $q$ in which it is at the moment.  Drawing a Moore machine we
658: indicate $h$ by labelling the states of its underlying finite
659: automaton by their values under $h.$ In addition, we almost always
660: make certain graphical simplifications: we merge all the transitions
661: joining the same pair of states into a single transition, labelled by
662: the union (evaluated in $\Sigma$) of all the labels. Sometimes we go
663: even farther and drop the label altogether from one transition, which
664: means that all the remaining input letters follow this transition.
665: 
666: Formally, a Moore machine computes a function $f_\A:\Omega^+\to
667: \Delta^+$ defined by
668: 
669: \[
670: f_\A(\omega_1\omega_2\dots\omega_n)
671: =h(\hat{\delta}(q_0,\omega_1))h(\hat{\delta}(q_0,\omega_1\omega_2))\dots
672: h(\hat{\delta}(q_0,\omega_1\omega_2\dots\omega_n))
673: \]
674: 
675: 
676: (note that $|f_\A(\omega_1\omega_2\dots\omega_n)|=n,$ as desired), and
677: a function $g_\A:\Omega^\infty\to \Delta^\infty$ defined by
678: 
679: \[g_\A(\omega_1\omega_2\dots)=
680: \bigsqcup\{f_\A(\omega_1\omega_2\dots\omega_n)~/~n=1,2,\dots\}.\]
681: 
682: 
683: 
684: We will be interested in Moore machines which compute $\trzy $-valued
685: functions.  This amounts to partitioning the state set $Q$ of $\A$
686: into three subsets $T,F,B,$ which we often make into parts of the
687: machine.  If we do so, we call the states in $T$ the {\em accepting
688: states\/} and the states in $F$ the {\em rejecting states}.  There
689: will be no special name for the states in $B.$
690: 
691: 
692: A Moore machine $\A$ is called {\em counter-free\/} if there is no
693: word $w\in\Omega^+$ and no states $q_1,q_2,\dots,q_s,\ s>1,$ such that
694: $\hat\delta(q_1,w)=q_2,\dots,\hat\delta(q_{s-1},w)=
695: q_s,\hat\delta(q_s,w)=q_1.$
696: 
697: 
698: 
699: \subsection{Markov chains}
700: 
701: For us, Markov chains are a synonym of {\em Markov chains with
702: stationary transitions and finite state space.}
703: 
704: 
705: Formally, given a finite set $I$ of {\em states} and a fixed
706: function $p:I\times I\to[0,1]$ satisfying 
707: 
708: \begin{equation}\label{stoch}
709: (\forall i\in I)\qquad\sum_{j\in I}p(i,j)=1,
710: \end{equation}
711: the {\em Markov chain\/} with state space $I$ and transitions $p$ is a
712: sequence $\X=X_0,X_1,\dots$ of random variables $X_n:W\to I$, such
713: that
714: 
715: \begin{equation}\label{M1}\Pr(X_{n+1}=j|X_n=i)=p(i,j).\end{equation}
716: 
717: The standard result of probability theory is that there exists a probability
718: triple $(W,\fM,\Pr)$ and a sequence $\X$ such that \eqref{M1} is
719: satisfied. $W$ is indeed the space of infinite sequences of ordered pairs
720: of elements from $I,$ and $\Pr$ is a certain product measure on this set.
721: 
722: 
723: One can arrange the values $p(i,j)$ in a matrix $\Pi=(p(i,j);i,j\in
724: I).$ Of course, $p(i,j)\geq 0$ and $\sum_{j\in I}p(i,j)=1$ for every
725: $i.$ Every real square matrix $\Pi$ satisfying these conditions is
726: called {\em stochastic.}  Likewise, the initial distribution of $\X$
727: is that of $X_0,$ which can be conveniently represented by a vector
728: $\Xi_0=(p(i); i\in I).$ Its choice is independent from the function
729: $p(i,j).$
730: 
731: It is often very convenient to represent Markov chains by matrices,
732: since many manipulations on Markov chains correspond to natural
733: algebraic operations performed on the matrices.
734: 
735: 
736: For our purposes, it is convenient to imagine the Markov chain $\X$ in
737: another, equivalent form: Let $K_I$ be the complete directed graph on
738: the vertex set $I.$ First we randomly choose the starting vertex in
739: $I,$ according to the initial distribution.  Next, we start walking in
740: $K_I;$ at each step, if we are in the vertex $i,$ we choose the edge
741: $(i,j)$ to follow with probability $p(i,j).$ If we define
742: $X_n=(\text{the vertex in which we are after $n$ steps}),$ then $X_n$
743: is indeed the same $X_n$ as in \eqref{M1}.
744: 
745: So we will be able to {\em draw\/} Markov chains.  Doing so, we will
746: often omit edges $(i,j)$ with $p(i,j)=0.$
747: 
748: \paragraph{Classification of states}
749: 
750: For two states $i,j$ of a Markov chain $\X$ with transition
751: probabilities $p$ we say that $i$ {\em communicates\/} with $j$ iff
752: there is a nonzero probability of eventually getting from $i$ to $j.$
753: Equivalently, it means that there is a sequence
754: $i=i_1,i_2,\dots,i_n=j$ of states such that $p(i_k,i_{k+1})>0$ for
755: $k=1,\dots,n-1.$ The reflexive relation of mutual communication (i.e.,
756: that $i$ communicates with $j$ and $j$ communicates with $i$ or $i=j$)
757: is an equivalence relation on $I.$ Class $[i]$ communicates with class
758: $[j]$ iff $i$ communicates with $j.$
759: 
760: The relation of communication is a partial ordering relation on
761: classes. The minimal elements in this partial ordering are called {\em
762: ergodic sets}, and nonminimal elements are called {\em transient
763: sets}. The elements of ergodic and transient sets are called ergodic
764: and transient states, respectively.
765: 
766: A Markov chain all whose ergodic sets are one-element is called {\em
767: absorbing}, and its ergodic states are called absorbing.
768: 
769: For ergodic sets one can be further define their {\em period.} Period
770: of an ergodic state $i$ is the gcd of all the numbers $p$ such that
771: there is a sequence $i=i_1,i_2,\dots,i_p=i$ of states such that
772: $p(i_k,i_{k+1})>0$ for $k=1,\dots,p-1.$ It can be shown that period is
773: a class property, i.e., all states in one ergodic class have the same
774: period. 
775: 
776: An ergodic set is called {\em aperiodic\/} iff its period is $1.$
777: Equivalently, it means that for every two $i,j$ in this set and all
778: sufficiently large $n$ there exists a sequence $i=i_1,i_2,\dots,i_n=j$
779: of states such that $p(i_k,i_{k+1})>0$ for $k=1,\dots,n-1.$
780: 
781: Every periodic class $C$ of period $p>1$ can be partitioned into $p$
782: periodic sub-classes $C_1,\dots,C_p$ such that $\Pr(X_{n+1}\in
783: C_{k+1\pmod p}|X_{n}\in C_{k\pmod p})=1$ for all $k.$
784: 
785: 
786: \section{Constructing conditionals}
787: 
788: We make a terminological distinction. If we speak about a {\em
789: conditional object}, we do not assume any probability space structure
790: imposed on $\Omega.$ When we have such structure
791: $(\Omega,\Sigma,\Pr),$ we speak about a {\em conditional event},
792: instead.
793: 
794: \subsection{Conditional objects}
795: 
796: First of all, let us note that any $\TL$ formula can be understood as
797: a definition of a pre-conditional from $\PC,$ which is indeed
798: $\dwa$-valued. Indeed, states of any model of temporal logic can be
799: interpreted as elements of $\Omega,$ and the whole model is thus an
800: element of $\Omega^+.$ The value the pre-conditional assigns to 
801: model $\M$ is $1$ if $\M,n\models\varphi$ and $0$ otherwise.
802: 
803: We construct a three-valued extension $\TT$ of $\TL$ as the set of
804: all pairs $(\varphi|\psi)$ of formulas from $\TL.$ The operator
805: $(\cdot|\cdot)$ can be understood as a {\em present tense
806: connective\/} of pre-conditionals, and, since formulas of $\TL$ are
807: $\dwa$-valued, it is sufficient to define its action as follows:
808: 
809: \[\begin{array}{|c|c|c|c|} \hline
810: \multicolumn{4}{|c|}{(x | y)}\\ \hline\hline x\diagdown y &0 &1 &\bot\\
811: \hline 0 &\bot &0 &~\\ \hline 1 &\bot &1 &~\\ \hline \bot &~ &~
812: &~\\ \hline
813: \end{array}
814: \]
815: 
816: \begin{definition}
817: A {\em conditional object of type 1\/} is a pre-conditional $c\in
818: \PC,$ definable in $\TT.$ The set of such conditional objects is denoted
819: $\CC.$
820: \end{definition}
821: 
822: 
823: \begin{definition} A {\em conditional object of type 2\/} is a pre-conditional
824: $c_+\in \PC_+,$ such that $c_+$ is computable by a $\trzy $-valued
825: counter-free Moore machine. The set of such conditional objects is denoted
826: $\CC_+.$
827: \end{definition}
828: 
829: \begin{definition} A {\em conditional object of type 3\/} is a pre-conditional
830: $c_\infty\in \PC_\infty,$ such that $c_\infty$ is computable by a $\trzy
831: $-valued counter-free Moore machine. The set of such conditional objects is
832: denoted $\CC_\infty.$
833: \end{definition}
834: 
835: 
836: The following proposition says that the conditional objects of types
837: 1, 2 and 3 are identical up to the way of representing
838: pre-conditionals.
839: 
840: \begin{theorem}\label{tfae}
841: \begin{align*}
842: \CC_+&=\{c_+\in \PC_+~/~c\in \CC\},\\
843: \CC_\infty&=\{c_\infty\in\PC_\infty~/~c\in\CC\},\\
844: \CC&=\{c\in \PC~/~c_\infty\in \CC_\infty\}.
845: \end{align*}
846: \end{theorem}
847: \begin{proof}
848: The equalities $\CC_+=\{c_+\in \PC_+~/~c_\infty\in \CC_\infty\}$ and
849: $\CC_\infty=\{c_\infty\in\PC_\infty~/~c_+\in\CC_+\}$ are obvious.
850: What remains to be proven are $\CC_+=\{c_+\in \PC_+~/~c\in \CC\}$ and
851: $\CC=\{c\in\PC~/~c_+\in\CC_+\}$
852: 
853: It is well-known \cite{Emerson} that propositional temporal logic of
854: past tense and (finite) deterministic automata are of equal expressive
855: power, i.e., in our terminology, the sets of $\dwa$-valued
856: pre-conditionals from $\PC$ definable in $\TL$ and computable by
857: deterministic finite automata are equal. Indeed the translations
858: between temporal logic and automata are effective.
859: 
860: We start with the first equality. Let $c$ be defined by a $\TT$
861: formula $(\varphi|\psi).$ Let $\A=(Q_\A,\Omega,\delta_\A,q_\A,T_\A)$
862: and $\B=(Q_\B,\Omega,\delta_\B,q_\B,T_\B)$ be deterministic finite
863: automata, computing the functions $\Omega^+\to\dwa$ defined by
864: $\varphi$ and $\psi,$ respectively.
865: 
866: Consider the Moore machine $(\A|\B)=(Q_\A\times
867: Q_\B,\Omega,\trzy,\delta,h,(q_\A,q_\B)),$ where
868: \begin{align*}
869: \delta((p,q),\omega)&=(\delta_\A(p,\omega),\delta_\B(q,\omega)),\\
870: h((p,q))&=\begin{cases}1&\text{if $p\in T_A$ and $q\in T_\B$},\\
871:                      0&\text{if $p\notin T_A$ and $q\in T_\B$},\\
872:                      \bot&\text{otherwise.}
873:         \end{cases}
874: \end{align*}
875: 
876: It is immediate to see that $(\A|\B)$ computes exactly
877: $(\varphi|\psi)_+.$
878: 
879: To prove the second equality, let $\A=(Q,\Omega,\trzy,\delta,h,q_0)$
880: be a Moore machine computing $c_+.$ We construct two deterministic
881: finite automata $\A_1=(Q,\Omega,\delta,q_0,\vec{h}^{-1}(\{1\}))$ and
882: $\A_2=(Q,\Omega,\delta,q_0,\vec{h}^{-1}(\{0,1\})$ from $\A,$ where
883: $\vec{h}^{-1}$ stands for the co-image under $h.$ Now let $\varphi_1$
884: and $\varphi_2$ be $\TL$ formulae corresponding to $\A_1$ and $\A_2,$
885: respectively. 
886: 
887: It is again immediate to see that $(\varphi_1|\varphi_2)$ defines
888: exactly the conditional in $\CC$ computed in $\CC_+$ by $\A.$
889: \end{proof}
890: 
891: Consequently, we can freely choose between the three available
892: representations of conditional objects. Doing so, we regard $\TT$ to
893: be the {\em logic of conditional objects}, while Moore machines
894: represent their {\em machine representation}. All these
895: representations are equivalent, thanks to Theorem~\ref{tfae}.
896: 
897: 
898: The classes $\CC,\ \CC_+$ and $\CC_\infty$ represent the {\em
899: semantics\/} of conditional objects, and again we can freely choose
900: the particular kind of semantical objects, thanks to
901: \eqref{reprezentacje}.
902: 
903: As an example, the simple conditional $(a|b)\in \TT$ is computed by
904: the following Moore machine.
905: 
906: 
907: \begin{figure}[!hbtp]
908: 
909: \[\UseTips
910: \xymatrix @C=15mm @R=10mm
911: {
912: {}\save[]+<5mm,-5mm>*{}\ar[rd]\restore&&*+++[o][F]{1}
913: \ar@(ul,ur)[]^{{ba}}
914: \ar[dd]^{{ba^\C}}
915: \ar@/_9mm/[dl]_{{b^\C}}\\
916: &*+++[o][F]{\bot}
917: \ar@(dl,l)[]^{{b^\C}}
918: \ar[ur]_{{ba}}
919: \ar@/_9mm/[dr]_{{ba^\C}}\\
920: &&*+++[o][F]{0}
921: \ar@(ld,dr)[]_{{ba^\C}}
922: \ar@/_9mm/[uu]_{{ba}}
923: \ar[ul]_{{b^\C}}
924: }
925: \]
926: \caption{Moore machine representing  conditional object
927: $(a|b).$}\label{f1}
928: \end{figure}
929: 
930: 
931: The above Moore machine, as it is easily seen, acts exactly according to the
932: rule ``ignore $b^\C$'s, decide depending on the truth status of $a$ when $b$
933: appears''. So indeed it represents the repetitions of the experiment for
934: $(a|b)$ according to the ``bridge'' repetition rule {\em start history anew}.
935: 
936: 
937: 
938: \subsection{Conditional events}
939: 
940: We will be using the name {\em conditional events\/} to refer to
941: conditionals considered with a probability space in the background.
942: 
943: Let $(\Omega,\PP(\Omega),\Pr)$ be a probability space.
944: 
945: \begin{definition}[Conditional event] Let $c\in \CC$ be a conditional object
946: over $\Omega.$ Suppose $\Omega$ is endowed with a probability space
947: structure $(\Omega,\Sigma,\Pr).$ With $c$ we associate the sequence
948: $\Y=\Y(c)=Y_1,Y_2,\dots$ of random variables $\Omega^\infty\to\trzy ,$
949: defined by the formula
950: 
951: \begin{equation}\label{e1}
952: Y_n(w)=\text{$n$-th-letter-of}(c_\infty(w)),
953: \end{equation}
954: 
955: where $\Omega^\infty$ is considered with the product probability structure.
956: 
957: We call $\Y$ the {\em conditional event\/} associated with $c,$ and denote it
958: $\ce{c},$ while $Y_n$ is then denoted $\ce{c}_n.$ Note that we do not include
959: the probability space in the notation. It will be always clear what
960: $(\Omega,\PP(\Omega),\Pr)$ is.
961: \end{definition}
962: 
963: In particular, $\Pr(\ce{c}_n=\true)$ is the probability that at time
964: $n$ the conditional is true, $\Pr(\ce{c}_n=\false)$ is the probability
965: that at time $n$ the conditional is false, and $\Pr(\ce{c}=\bot)$ is
966: the probability that at time $n$ the conditional is undefined.
967: 
968: 
969: \begin{definition}[Probability of conditional events]~\\
970: We define the {\em asymptotic probability at time $n$\/} of a
971: conditional $c$ by the formula
972: 
973: \begin{equation}\label{e2}
974: \pr_n(c)=\dfrac{\Pr(\ce{c}_n=1)}
975: {\Pr(\ce{c}_n=0\ \text{or}\ 1)}.
976: \end{equation}
977: 
978: If the denominator is $0,$ $\pr_n(c)$ is undefined.
979: 
980: The {\em asymptotic probability\/} of $c$ is
981: 
982: \begin{equation}\label{e3.5}
983: \Pr(c)=\lim_{n\to\infty}\pr_n(c),
984: \end{equation}
985: 
986: provided that $\pr_n(c)$ is defined for all sufficiently large $n$ and
987: the limit exists.
988: 
989: We will regard $\ce{c}$ as {\em probabilistic semantics\/} of $c.$
990: 
991: If $\varphi\in\TL$ then we write $\Pr(\varphi)$ for $\Pr((\varphi|\TRUE)).$
992: \end{definition}
993: 
994: 
995: It is perhaps reasonable to explain why we want the conditional event
996: and its probability to be defined in this way.  The main motivation is
997: that we want the conditional event and its probability to be natural
998: and intuitive. And we achieve this by using the recipe of de Finetti,
999: which in our case materializes in the above definitions.
1000: 
1001: 
1002: 
1003: \section{Underlying Markov chains, Bayes' Formula and classification of
1004: conditional events} 
1005: \label{classifying}
1006: 
1007: 
1008: \subsection{Underlying Markov chains}
1009: 
1010: Let $c$ be a conditional object and let $\A=(Q,\Omega,\delta,\trzy,h,q_0)$ be
1011: a counter-free Moore machine which computes $c_\infty.$
1012: 
1013: We define a Markov chain $\X=\X(\A)$ by taking the set of states of
1014: $\X$ to be the set $Q$ of states of $\A,$ and the transition function
1015: $p$ to be defined by
1016: 
1017: \begin{equation*}
1018: p(q,q')=\sum_{\substack{\omega\in\Omega\\\delta(q,\omega)=q'}}\Pr(\{\omega\}).
1019: \end{equation*}
1020: 
1021: Indeed, for every $q$ we have
1022: 
1023: \begin{equation*}
1024: \sum_{q'}p(q,q')
1025: =\sum_{q'}\sum_{\substack{\omega\in\Omega\\
1026: \delta(q,\omega)=q'}}\Pr(\{\omega\}) 
1027: =\sum_{\omega\in\Omega}\Pr(\{\omega\}) =1,
1028: \end{equation*}
1029: 
1030: which means that the function $p$ satisfies \eqref{stoch}, which is
1031: the criterion for being a transition probability function of a Markov
1032: chain.  The initial probability distribution is defined by
1033: 
1034: \[
1035: p(q)=
1036: \begin{cases}1&\text{if $q=q_0,$ the initial state of $\A,$}\\
1037:              0&\text{otherwise.}
1038: \end{cases}
1039: \]
1040: 
1041: Therefore we have indeed converted $\A$ into a Markov chain $\X.$
1042: 
1043: 
1044: In the pictorial representation of the conversion process is much
1045: simpler: we take the drawing of $\A,$ and replace all the letters from
1046: $\Omega$ marking transitions by their probabilities according to $\Pr,$
1047: and then contract multiple transitions between the same states into a
1048: single one, summing up their probabilities.
1049: 
1050: \begin{theorem}\label{transient-aperiodic} $\X$ is a Markov chain in
1051: which only transient and aperiodic states exist.
1052: \end{theorem}
1053: 
1054: \begin{proof} Suppose $\X$ has a periodic set $C$ of period $p>1,$ and
1055: $C_1,\dots,C_p$ its division into periodic subclasses. Let $\omega\in
1056: \Omega$ be any atomic event with $\Pr(\{\omega\})>0.$ Let $q\in C_1.$
1057: Since $\Pr(X_{n+1}\in C_{k+1\pmod p}|X_{n}\in C_{k\pmod p})=1$ for all
1058: $k,$ it follows that $\delta^1(q,\omega)=\delta(q,\omega)\in C_{2\pmod
1059: p},$ and likewise
1060: $\delta^{k+1}(q,\omega)=\delta(\delta^k(q,\omega))\in C_{k+1\pmod p}$
1061: for $k\geq 1.$
1062: 
1063: However, $C$ is finite, so there must be $s\neq t$ such that
1064: $\delta^s(q,\omega)=\delta^t(\omega).$
1065: 
1066: The sequence 
1067: 
1068: \[\delta^s(q,\omega),\delta^{s+1}(q,\omega),\dots,\delta^t(q,\omega)
1069: =\delta^s(q,\omega)\] 
1070: 
1071: thus violates the assumption that $\A$ is counter-free.\end{proof}
1072: 
1073: The next corollary follows by the classical result about finite Markov chains.
1074: 
1075: 
1076: \begin{corollary}\label{limits}
1077: For every state $i$ of $\X,$ the limit $\lim_{n\to\infty}\Pr(X_n=i)$
1078: exists.
1079: \end{corollary}
1080: 
1081: 
1082: 
1083: Using $h:Q\to\trzy,$ the {\em acceptance mapping of $\A,$} we get
1084: 
1085: \begin{theorem}
1086: \(\label{Y=h(X)}\ce{c}=h(\X).\)\qed
1087: \end{theorem}
1088: 
1089: 
1090: Note that $\ce{c}$ defined above need not be a Markov chain
1091: itself, but it is a simple projection of a Markov chain, extracting
1092: all the invariant information.  Of course, it will be typically very
1093: beneficial to work most of the time with $\X,$ having the whole theory
1094: of Markov chains as a tool-set, and only then to move to $\ce{c}.$
1095: 
1096: Let us examine the previously given definition of $(a|b)$ to see what its
1097: probability is.
1098: 
1099: The Markov chain looks as follows: 
1100: 
1101: 
1102: \begin{figure}[!hbtp]
1103: 
1104: \[\UseTips
1105: \xymatrix @C=15mm @R=10mm
1106: {
1107: {}\save[]+<5mm,-5mm>*{}\ar[rd]\restore
1108: &&*+++[o][F]{1}
1109: \ar@(ul,ur)[]^{{\Pr(ba)}}
1110: \ar[dd]^{{\Pr(ba^\C)}}
1111: \ar@/_9mm/[dl]_{{\Pr(b^\C)}}\\
1112: &*+++[o][F]{\bot}
1113: \ar@(dl,l)[]^{{\Pr(b^\C)}}
1114: \ar[ur]_{{\Pr(ba)}}
1115: \ar@/_9mm/[dr]_{{\Pr(ba^\C)}}\\
1116: &&*+++[o][F]{0}
1117: \ar@(ld,dr)[]_{{\Pr(ba^\C)}}
1118: \ar@/_14mm/[uu]_{{\Pr(ba)}}
1119: \ar[ul]_{{\Pr(b^\C)}}
1120: }
1121: \]
1122: \caption{Markov chain corresponding to the Moore machine on Fig.\
1123: \ref{f1}.}\label{f2}
1124: \end{figure}
1125: 
1126: where the initial distribution assumes probability 1 given to the
1127: state pointed to by the arrow ``from nowhere''.  
1128: 
1129: It is easy to check that $\Pr((a|b))=\Pr(ba)/\Pr(b),$ provided that
1130: $\Pr(b)>0.$ Indeed, for every $n$ holds $\Pr(\ce{(a|b)}_n=1)=\Pr(ba)$
1131: and $\Pr(\ce{(a|b)}_n=0)=\Pr(ba^\C),$ so $\Pr(\ce{(a|b)}_n=0\
1132: \text{or}\ 1)=\Pr(ba)+\Pr(ba^\C)=\Pr(b).$ It is so because, no matter
1133: in which state we are, these are the probabilities of getting to $1$
1134: and $0$ in the next step, respectively. This evaluation will follow from
1135: Bayes' Formula below, too.
1136: 
1137: \subsection{Bayes' Formula}
1138: First of all, let us note that for each $\star\in\trzy$ the limit
1139: $\lim_{n\to\infty}\pr(\ce{c}_n=\star)$ exists, since, for any choice
1140: of a Moore machine $\A$ computing $c_+$ and assuming $\X=\X(\A),$
1141: $\pr(\ce{c}_n=\star)$ is a sum of $\pr(X_n=i)$ over all states $i$ of
1142: $\X$ with $h(i)=\star,$ and the latter probabilities converge by
1143: Corollary \ref{limits}.
1144: 
1145: A conditional event is called {\em regular\/} iff
1146: $\lim_{n\to\infty}\Pr(\ce{c}_n=0\ \text{or}\ 1)>0.$ In particular, for
1147: regular conditionals the limit in \eqref{e3.5} always exists and is
1148: equal to
1149: 
1150: \begin{equation*}
1151: \dfrac{\lim_{n\to\infty}\Pr(\ce{c}_n=1)}
1152: {\lim_{n\to\infty}\Pr(\ce{c}_n=0\ \text{or}\ 1)}.
1153: \end{equation*}
1154: 
1155: Turning to the logical representation of conditionals,  we have thus
1156: 
1157: \begin{theorem}[Bayes' Formula]\label{Bayes}
1158: For $(\varphi|\psi)\in\TT$ 
1159: \[\Pr((\varphi|\psi))=\frac{\Pr(\varphi\land\psi)}{\Pr(\psi)}\]
1160: 
1161: whenever the right-hand-side above is well-defined.\qed
1162: \end{theorem}
1163: 
1164: Note that Bayes' Formula has been expected by de Finetti for the frequency
1165: based conditionals.
1166: 
1167: \subsection{Classifying conditional events}
1168: 
1169: It is interesting to consider the conditionals $c$ for which
1170: $\lim_{n\to\infty}\Pr(\ce{c}_n=0\ \text{or}\ 1)=0.$ We can distinguish
1171: two types of such conditional events: those for which $\Pr(\ce{c}_n=0\
1172: \text{or}\ 1)$ is identically $0$ for infinitely many $n,$ and those
1173: for which it is nonzero for all but finitely many $n$. The former will
1174: be called {\em degenerate}, the latter {\em strange}. We call {\em
1175: strictly degenerate\/} those degenerate events, for which
1176: $\Pr(\ce{c}_n=0\ \text{or}\ 1)$ for all but finitely many $n.$
1177: 
1178: The degenerate conditional events correspond to bets which infinitely
1179: often cannot be resolved, because they are undefined, and strictly
1180: degenerate events are those which are almost never defined.
1181: 
1182: Strange conditional events are more interesting. The Bayes' Formula is
1183: senseless for them, so we have to use some ad hoc methods to see if
1184: their asymptotic probability exists or not.
1185: 
1186: The first example shows that the sequence $\pr_n(c)$ can be
1187: nonconvergent for strange $c.$
1188: 
1189: Consider $c_1=(a|\PBOX ((\PREV a\to a^\C)\land(\PREV a^\C\to
1190: a)\land(\lnot\PREV \TRUE\to a ))),$ where $0<\Pr(a)<1.$ The long
1191: temporal formula asserts that $a$ always follows $a^\C$ and $a^\C$
1192: always follows $a,$ and at the beginning of the process ($n=1$), where
1193: $\PREV \TRUE $ is false, $a$ holds.
1194: 
1195: It is easily verified that 
1196: 
1197: \[\pr_n(\ce{c_1})=\begin{cases}
1198:                  0&\text{if $n$ is even},\\
1199:                  1&\text{if $n$ is odd}.
1200:            \end{cases}
1201: \]
1202: 
1203: Thus the finite-time behaviour of this conditional is not
1204: probabilistic---its truth value depends solely on the {\em age\/} of
1205: the system. So for somebody expecting a pure game of chances its
1206: behaviour must seem strange (and hence the name of this class of
1207: conditional events).
1208: 
1209: Note that we have just discovered the next feature of conditionals
1210: expected by de Finetti: nonconvergence of the limiting frequency when
1211: probability of the `given' part tends to $0.$
1212: 
1213: However, again following de Finetti, if $(\varphi|\varphi)$ is
1214: strange, its asymptotic probability is $1.$ E.g., $\Pr(\PBOX ((\PREV
1215: a\to a^\C)\land(\PREV a^\C\to a)\land(\lnot\PREV \TRUE\to a ))|\PBOX
1216: ((\PREV a\to a^\C)\land(\PREV a^\C\to a)\land(\lnot\PREV \TRUE\to a
1217: )))=1.$
1218: 
1219: Moreover, for $c_2=(a|\PBOX ((\PREV a\to  a^\C)\land(\PREV a^\C\to 
1220: a)))$ we have 
1221: \[\pr_n(\ce{c_2})=\begin{cases}
1222:                  1-\Pr(a)&\text{if $n$ is even},\\
1223:                  \Pr(a)&\text{if $n$ is odd}.
1224:            \end{cases}
1225: \]
1226: 
1227: Indeed, here the `given' part requires that $a$'a and $a^\C$'s
1228: alternate, but does not specify what is the case at the beginning of
1229: the process. So the probability of the whole conditional at odd times
1230: is the probability that $a$ has happened at time 1, and at even times
1231: it is the probability that $a$ has not happened at time 1.  Therefore,
1232: when $\Pr(a)=1/2,$ $\Pr(c_2)$ exists and is $1/2.$ So asymptotic
1233: probabilities which are neither $0$ nor $1$ are possible for strange
1234: conditionals events.
1235: 
1236: At present, the question whether there it is decidable if a given
1237: strange conditional event has an asymptotivc probability is open.
1238: However, we believe that te answer is positive and offer it as our
1239: cojecture.
1240: 
1241: 
1242: \begin{conj}
1243: The set of conditional events which have asymptotic probability is
1244: decidable.  Moreover, for those events which have asymptotic
1245: probability, its value is effectively computable. 
1246: \end{conj}
1247: 
1248: 
1249: 
1250: 
1251: 
1252: \section{Connectives of conditionals}
1253: 
1254: \subsection{Present tense connectives}
1255: 
1256: Let us recall that present tense connectives are those, whose definition
1257: in $\TT$ does not use temporal connectives, and therefore depends
1258: on the present, only. Equivalently, an $n$-ary present tense connective is
1259: completely characterised by a function $\trzy ^n\to\trzy .$ 
1260: 
1261: Here are several possible choices for the conjunction, which is always
1262: defined as a pointwise application of the following $\trzy $ valued
1263: functions. Above we display the notation for the corresponding kind of
1264: conjunction.
1265: 
1266: \[
1267: \conn{\land_{\SAC}}00001101\ \ \ \ 
1268: \conn{\land_{\GNW}}00001\bot0\bot\ \ \ \
1269: \conn{\land_{\Sch}}00\bot01\bot\bot\bot
1270: \]
1271: \[{\begin{array}{|c|c|}\hline 
1272: \multicolumn{2}{|c|}{\sim x} \\\hline\hline x&\sim x\\\hline 0&1\\\hline
1273: 1&0\\\hline\bot&\bot\\\hline\end{array}}
1274: \]
1275: \[\conn{\lor_{\SAC}}01011101\ \ \ \ 
1276: \conn{\lor_{\GNW}}0111\bot1\bot\bot\ \ \ \
1277: \conn{\lor_{\Sch}}01\bot11\bot\bot\bot.
1278: \]
1279: 
1280: They can be equivalently described by syntactical manipulations in $\TT.$
1281: The reduction rules are as follows:
1282: 
1283: \begin{equation}\label{TLTL}
1284: \begin{split}
1285: (a|b)\land_\SAC(c|d)&=(abcd\lor abd^\C\lor cdb^\C|b\lor d)\\
1286: (a|b)\land_\GNW(c|d)&=(abcd|a^\C d\lor c^\C d\lor abcd)\\
1287: (a|b)\land_\Sch(c|d)&=(abcd|bd)\\
1288: \sim (a|b)&=(a^\C|b)\\
1289: (a|b)\lor_\SAC(c|d)&=(ab\lor cd|b\lor d)\\
1290: (a|b)\lor_\GNW(c|d)&=(ab\lor cd|ab\lor cd \lor bd)\\
1291: (a|b)\lor_\Sch(c|d)&=(ab\lor cd|bd).
1292: \end{split}
1293: \end{equation}
1294: 
1295: 
1296: The first is based on the principle ``if any of the arguments becomes
1297: defined, act!''.  A good example would be a quotation from \cite{c97}:
1298: 
1299: \begin{quote}\sl ``One of the most dramatic examples of the unrecognised
1300: use of compound conditioning was the first military strategy of our
1301: nation.   As the Colonialists waited for the British to attack, the
1302: signal was `One if by land and two if by sea'.  This is the
1303: conjunction of two conditionals with uncertainty!''
1304: \end{quote}
1305: 
1306: Of course, if the above was understood as a conjunction of two
1307: conditionals, the situation was crying for the use of $\land_\SAC,$
1308: whose definition has been proposed independently by Schay, Adams and
1309: Calabrese (the author of the quotation).
1310: 
1311: The conjunction $\land_{\GNW}$ represents a moderate approach, which
1312: in case of an apparent evidence for $\false$ reports $\false,$ but
1313: otherwise it prefers to report unknown in a case of any doubt. Note
1314: that this conjunction is essentially the same as {\em lazy
1315: evaluation},  known from programming languages. 
1316: 
1317: Finally, the conjunction $\land_\Sch$ is least defined, and acts
1318: (classically) only if both arguments become defined. It corresponds to
1319: the {\em strict evaluation}.
1320: 
1321: We have given an example for the use of $\land_\SAC.$ The uses of
1322: $\land_\GNW$ and $\land_\Sch$ can be found in any computer program
1323: executed in parallel, which uses either lazy or strict evaluation of
1324: its logical conditions. And indeed both of them happily coexist in
1325: many programming languages, in that one of them is the standard
1326: choice, the programmer can however explicitly override the default and
1327: choose the other evaluation strategy.
1328: 
1329: Let us mention that all the three systems above are in fact
1330: well-known, classical so to say three-valued logics: 
1331: $\langle\land_\GNW,\lor_\GNW,\sim\rangle$ is the  logic of
1332: \L{}ukasiewicz, $\langle\land_\SAC,\lor_\SAC,\sim\rangle$ is the logic
1333: of Soboci\'nski, and $\langle\land_\Sch,\lor_\Sch,\sim\rangle$ is the
1334: logic of Bochvar.
1335: 
1336: \subsection{Past tense connectives} 
1337: 
1338: The following connective is tightly related to very close to the conjunction
1339: of the {\em product space conditional event algebra\/} introduced in
1340: \cite{g94}. Detailed discussion of embeddings of existing algebras of
1341: conditional events into $\TT$ is included in the companion paper
1342: \cite{cea}. Our new conjunction, denoted $\land^\star,$ is defined
1343: precisely when at least one of its arguments is defined, so it
1344: resembles $\land_\SAC$ in this respect, but instead of assigning the
1345: other argument a default value when it is undefined, like $\SAC$ does,
1346: it uses its most recent defined value, instead.  However, when the
1347: other argument hasn't ever been defined, it is assumed to act like
1348: $\FALSE.$
1349: 
1350: In the language of $\TT$ $(a|b)\land^\star(c|d)$ can be expressed by
1351: \[((b^\C\Since (a\land b))\land(d^\C\Since (c\land
1352: d))|b\lor d).\]
1353: 
1354: 
1355: 
1356: 
1357: \subsection{Conclusion}
1358: 
1359: We believe that there is no reason to restrict our attention to any
1360: particular choice of an operation extending the classical conjunction,
1361: and call is {\em the conjunction of conditionals.} There are indeed
1362: many reasonable such extensions, which correspond to different
1363: intuitions and situations, they can coexist in a single formalism, and
1364: any restriction in this respect necessarily narrows the applicability
1365: of the formalism.
1366: 
1367: We believe that neither of the choices discussed in this paragraph is {\em
1368: the\/} conjunction of conditionals. There are indeed many possible choices,
1369: and all of them have their own merits.  In fact already the original system of
1370: Schay consisted of five operations:  $\sim,\land_\SAC,\lor_\SAC,\land_\Sch$
1371: and $\lor_\Sch.$ Moreover, he was aware that these operations still do not
1372: make the algebra functionally complete (even in the narrowed sense, restricted
1373: to defining only operations which are undefined for all undefined arguments).
1374: And in order to remedy this he suggested to use one of several additional
1375: operators, one of them being $\land_\GNW!$ So for him all those operations
1376: could coexist in one system.
1377: 
1378: \section{Three prisoner's puzzle}
1379: 
1380: 
1381: In order to demonstrate that our formalism allows for a precise
1382: treatment of problems with conditioning and probabilities, let us
1383: consider the following classical example of a probabilistic
1384: ``paradox''.  We will take this opportunity to highlight some of the
1385: practical issues of modelling using $\TT$ and Moore machines
1386: approach. Therefore our analysis will be very detailed.
1387: 
1388: \subsection{The puzzle}
1389: 
1390: The three prisoner's puzzle  \cite{p88} is the following: 
1391: 
1392: \begin{quote}
1393: Three prisoners are sentenced for execution. One day before their
1394: scheduled execution, prisoner $A$ learns that two of them have been
1395: pardoned.  $A$ calculates a probability of $2/3$ for him being
1396: pardoned.  Then he asks the Guard: ``Name me one of my fellows who
1397: will be pardoned.  The Guard tells him, that $B$ will be pardoned.
1398: Based on that information, $A$ recalculates the probability of being
1399: pardoned as $1/2$, since now only one pardon remains for him and $C$
1400: (the third prisoner) to share!  However, he could apply the same
1401: argument if the Guard had named $C$.  Furthermore, he knew beforehand
1402: that at least one of his fellows will be pardoned --- so what did he
1403: gain (or lose) by the answer?
1404: \end{quote}
1405: 
1406: The intuitive explanation is that after learning the Guard's testimony
1407: $G(B)$ that $B$ will be pardoned, $A$ should revise the probability of
1408: the event $P(A)$ (of him being pardoned) by computing $P(\PREV
1409: P(A)|G(B))$, and the probability evaluation yields in this case
1410: $2/3,$ as expected.
1411: 
1412: However, what he indeed calculated was $P(\PREV G(B)|P(A))$, assuming
1413: effectively that the pardon had been given with equal probabilities to
1414: all the pairs possible {\em after\/} Guard's testimony. This
1415: probability turns out to be $1/2.$
1416: 
1417: \subsection{Probability tree model}
1418: 
1419: First we present a simple probability tree analysis of the paradox,
1420: using the method which originates with Huygens \cite{huygens,shafer}
1421: and is indeed almost as old as the mathematically rigorous probability
1422: theory itself. We begin in the leftmost circle (before pardon), then
1423: each of the three pardoned pairs leads us to three next circles,
1424: indicating the situation after the pardon. Finally, we have all the
1425: possible testimonies of the Guard. All edges originating from the same
1426: circle are equiprobable. After Guard's testimony $G(B),$ only the two
1427: top circles on the right are possible, and their probabilities are in
1428: the proportion $2:1,$ the more probable one being the one in which $A$
1429: is pardoned, while he is executed in the other one. So indeed even
1430: after the testimony the probability that $A$ is pardoned remains
1431: $2/3.$ 
1432: \begin{figure}[!hbtp]
1433: \[\UseTips
1434: \xymatrix @C=20mm
1435: {
1436: &&*+++[o][F]{}
1437: \\
1438: &*+++[o][F]{}
1439: \ar[ur]^{\textstyle{G(B)}}
1440: &*+++[o][F]{}
1441: \\
1442: *+++[o][F]{}
1443: \ar[ur]^{\textstyle{P(AB)}}
1444: \ar[r]^(.7){\textstyle{P(BC)}}
1445: \ar[dr]_{\textstyle{AC}}
1446: &*+++[o][F]{}
1447: \ar[ur]^{\textstyle{G(B)}}
1448: \ar[dr]_{\textstyle{G(C)}}
1449: \\
1450: &*+++[o][F]{}
1451: \ar[dr]_{\textstyle{G(C)}}
1452: &*+++[o][F]{}
1453: \\
1454: &&*+++[o][F]{}
1455: }
1456: \]
1457: \caption[]{Probability tree analysis of the three prisoner
1458: puzzle.}
1459: \label{f6}
1460: \end{figure}
1461: 
1462: \subsection{$\TT$ and Moore machine models}
1463: 
1464: However, the tree shown above strongly resembles a Moore machine. And
1465: indeed, we augment it with the necessary details below. The most
1466: substantial change is that the Moore machine requires the same set of
1467: atomic possibilities is given at each state, which determine the next
1468: transition. Therefore:
1469: 
1470: \begin{itemize}
1471: \item The Guard testifies something irrelevant while the court decides
1472: the pardons, and the court decides something irrelevant while the Guard
1473: testifies. This change is made invisible by our convention of
1474: collapsing transitions and applying subsequently Boolean algebra
1475: simplifications, except that
1476: \item In cases when the Guard has no choice, we must replace the
1477: existing transition label by the full event, because the Guard has
1478: prescribed answer no matter whom he would like to name,
1479: \item And except that we have to decide about transitions from the
1480: states which are terminal in the tree model.  Because we believe that
1481: after being pardoned nobody can be prosecuted again for the same
1482: crime, and we do not believe in reincarnation, either, our choice is
1483: to use self-loops in the terminal states, yielding a ``Russian
1484: roulette'' model.
1485: \end{itemize}
1486: 
1487: 
1488: \begin{figure}[!hbtp]
1489: \[\UseTips
1490: \xymatrix @C=20mm
1491: {
1492: &&&*+++[o][F]{\textcolor{green}{1}}
1493: \ar@{.>}@(ur,dr)[]
1494: \\
1495: &&*+++[o][F]{\textcolor{green}{\bot}}
1496: \ar[ur]^{\textcolor{green}{\textstyle{\Omega}}}
1497: &*+++[o][F]{\textcolor{green}{0}}
1498: \ar@(ur,dr)@{.>}[]
1499: \\
1500: \ar@{.>}[r]
1501: &*+++[o][F]{\textcolor{green}{\bot}}
1502: \ar[ur]^{\textstyle{P(AB)}}
1503: \ar[r]^(.67){\textstyle{P(BC)}}
1504: \ar[dr]_{\textstyle{AC}}
1505: &*+++[o][F]{\textcolor{green}{\bot}}
1506: \ar[ur]_{\textstyle{G(B)}}
1507: \ar[dr]^{\textstyle{G(C)}}\\
1508: &&*+++[o][F]{\textcolor{green}{\bot}}
1509: \ar[dr]_{\textcolor{green}{\textstyle{\Omega}}}
1510: &*+++[o][F]{\textcolor{green}{\bot}}
1511: \ar@{.>}@(ur,dr)[]
1512: \\
1513: &&&*+++[o][F]{\textcolor{green}{\bot}}
1514: \ar@{.>}@(ur,dr)[]
1515: }
1516: \]
1517: \caption[]{Probability tree analysis of the three prisoner puzzle with
1518: {\color{green} extensions necessary to convert the diagram into a Moore
1519: machine.}}
1520: \label{f7}
1521: \end{figure}
1522: 
1523: This provides a next piece of evidence that our definition of
1524: conditional events is natural and close to intuitions.  In fact, one
1525: can embed the whole probability tree model into the formalism of
1526: Russian roulette Markov chains \cite{shafer}, and thus shows that our
1527: model of conditionals extends the method of probability trees.
1528: 
1529: Next we attempt to model the same paradox syntactically in $\TT.$ The
1530: construction of a correct $\TT$ representation is a little bit more
1531: complicated than the formula $P(\PREV P(A)|G(B))$ we have suggested
1532: previously, as this requires specifying the actions of the Guard,
1533: whose probabilities are affected by the pardon decision. So we assume
1534: that the Guard always tosses a coin. If he gets heads $(H)$, he tells
1535: the alphabetically first name among those applicable, and in case of
1536: tails $(T)$ the alphabetically last among them. This indicates the
1537: need to consider the strategy followed by the Guard. And in fact, the
1538: probabilities $A$ calculates depend on what he assumes about this
1539: strategy. So indeed now the answers of the Guard are shorthands for
1540: the combinations of the pardon decision and the coin toss outcome.
1541: Therefore $G(B)$ is $(\PREV P(AB)\land(H\lor T))\lor (\PREV P(BC)\land
1542: H).$ 
1543: 
1544: 
1545: Moreover, we have to decide what should be modelled by the conditional
1546: object, and what by the probability assignment, which turns the former
1547: into a stochastic process.  The general rule is that the more of the
1548: modelling is encoded in the probability assignment, the simpler the
1549: conditional and its Moore machine are.  On the other hand, encoding
1550: everything in the probability distribution is difficult and prone to
1551: errors, as the example of the poor prisoner shows. An, needless to
1552: say, a good model is one in which the proportions are just right. More
1553: on that below.
1554: 
1555: So formally the conditional looks now as follows:
1556: 
1557: \begin{equation}\label{3pris}
1558: \left(
1559: (\PREV AB)
1560: \lor
1561: (\PREV AC)
1562: \left|
1563: ((\PREV AB)\land (H\lor T))
1564: \lor
1565: ((\PREV BC)\land H)
1566: \right.\right),
1567: \end{equation}
1568: 
1569: with $\E=\{P(AB),P(BC),AC,H,T\},$ where the events $P(AB),$ $P(BC)$
1570: and $AC$ mutually exclusive and equiprobable, and similarly $H$ and
1571: $T$ mutually exclusive and equiprobable. (Our construction will easily
1572: handle non-equal probabilities, i.e., biased pardon decision and/or
1573: biased coin, too.) So the set $\Omega$ of atomic events is $\{ AB\,H,
1574: AB\,T, BC\,H, BC\,T,AC\,H, AC\,T\}, $ and these events are
1575: equiprobable under our probability assignment. However, we will be
1576: able to calculate the probability of \eqref{3pris} without the
1577: equiprobability assumption, too.
1578: 
1579: Note that, e.g., assuming events $A,B$ and $C$ to be nonexclusive
1580: individual pardon decisions of probability $1/3$ each, leads to more
1581: complicated conditional expression, because a substantial amount of
1582: coding effort must used just to ensure that always precisely two
1583: prisoners are pardoned. This makes the Moore machine more complicated,
1584: too. So this is certainly not a good model, because what can be easily
1585: taken care of by the probability assignment is instead modelled by
1586: logical methods. Such a model can be of course
1587: correct,\footnote{Although unnecessary complications certainly
1588: increase the risk of mistakes and make verification of the model
1589: harder.} but good means for us more than just correct.
1590: 
1591: But if we attempt to draw the Moore machine of our conditional, we
1592: discover that it is quite different from that on Fig.\ \ref{f7}.
1593: 
1594: \begin{figure}[!hbtp]
1595: \[\UseTips
1596: \xymatrix @R=18mm @C=9mm
1597: {
1598: &&*+++[o][F]{\bot}
1599: &&*+++[o][F]{0}
1600: \\
1601: &*+++[o][F]{1}
1602: \ar@(l,u)^(.63){AB(H\lor T)}
1603: \ar[rrrr]|(.4){BC(H\lor T)}
1604: \ar[ddrr]|(.44){AC(H\lor T)}
1605: &&&&*+++[o][F]{1}
1606: \ar[llllld]|{AB\,H}
1607: \ar[lllu]|{AB\,T}
1608: \ar[ul]|{BC\,H}
1609: \ar[dr]|{BC\,T}
1610: \ar[dd]|(.33){AC\,T}
1611: \ar[lllldd]|(.375){AC\,H}[]
1612: \\
1613: *+++[o][F]{0}
1614: &&&&&&*+++[o][F]{\bot}
1615: \\
1616: &*+++[o][F]{0}
1617: \ar[uuur]|(.25){AB(H\lor T)}
1618: \ar[urrrrr]|{BC(H\lor T)}
1619: \ar@/_5ex/[rrrr]_{AC(H\lor T)}
1620: &&*+++[o][F]{1}
1621: &&*+++[o][F]{\bot}
1622: \\
1623: &&&&&
1624: \ar@{.>}[r]
1625: &*+++[o][F.]{\bot}
1626: \ar@{.>}[uu]^(.7){BC}
1627: \ar@{.>}@/_2.3em/[uuuullll]|(.19){AB}
1628: \ar@{.>}[ul]^{AC}
1629: }
1630: \]
1631: \caption[]{Moore machine corresponding to formula
1632: \eqref{3pris}.}\label{f9}
1633: \end{figure}
1634: 
1635: The overall structure of the Moore machine is as follows: The entry
1636: states and transitions are dotted. Each of the three lines of three
1637: states (they form roughly edges of a triangle), consists of states
1638: with the same, already known pardon decision in the next experiment,
1639: while the current experiment's outcome is represented as the label of
1640: the state. Transitions are shown for one state on each edge only,
1641: because their targets depend on the input only, and not on the source
1642: within that edge.  And this is why we can calculate the probability of
1643: \eqref{3pris} in a quite straightforward way. For time greater than
1644: $1$ the probability of getting in two steps to a state with a given
1645: label does not depend on the current state nor on the
1646: time. Essentially, after the first step the edge of the triangle is
1647: chosen, which corresponds to the move to one of the states in the
1648: middle column of Fig.\ \ref{f7}. In the second step we move to the
1649: state with the label equal to the destination label from Fig.\
1650: \ref{f7}, and the edge it is found within depends on the next
1651: experiment, already. The similarity is even stronger if we compare
1652: Fig. \ref{f9} with Fig.\ \ref{3prisRR} rather than with Fig.\
1653: \ref{f7}. A formal calculation, using matrix calculus, can be found in
1654: Section \ref{algorytmy} below.
1655: 
1656: The most substantial difference is that \eqref{3pris} is not a
1657: ``Russian roulette'' model! To note this set time to $3$ and see: the
1658: present outcomes depend on the pardon decisions made at time $2$,
1659: while the Guard was testifying in the previous round of the
1660: experiment, and while we are hearing the testimony of the Guard now,
1661: the pardons are already decided as a part of the next experiment. So
1662: the probabilistic choices which we described as irrelevant for the
1663: Moore machine model, are parts of the previous/next repetition schema
1664: here. The overlapping experiments do not interfere, however, so this
1665: does not affect probabilities.  Furthermore, all the final outcome
1666: undefined values have been merged into one state. Finally, there are
1667: entry states which are visited just once and correspond to the
1668: situation at time $1$, when the Guard says something, but there is no
1669: pardon decision to compare it with.
1670: 
1671: A modified version of \eqref{3pris}, which is Russian roulette, is as
1672: follows: 
1673: 
1674: \begin{equation}\label{3prisRR}
1675: \left(
1676: (\atone AB)
1677: \lor
1678: (\atone AC)
1679: \left|
1680: ((\atone AB)\land \attwo(H\lor T))
1681: \lor
1682: ((\atone BC)\land \attwo H)
1683: \right.\right),
1684: \end{equation}
1685: 
1686: where $\atone\alpha$ is $\PDIA(\lnot\PREV\TRUE\land \alpha)$ and
1687: $\attwo\alpha$ is $\PDIA(\PREV\TRUE\land \lnot\PREV\PREV\TRUE\land
1688: \alpha),$ and express that $\alpha$ is true at time $1$ and $2$,
1689: respectively.
1690: 
1691: 
1692: \begin{figure}[!hbtp]
1693: \[\UseTips
1694: \xymatrix @C=20mm
1695: {
1696: &&&*+++[o][F]{1}
1697: \ar@(ur,dr)[]
1698: \\
1699: &&*+++[o][F]{{\bot}}
1700: \ar[ur]^{\textstyle{H\lor T}}
1701: &*+++[o][F]{{0}}
1702: { \ar@(ur,dr)[]}
1703: \\
1704: {\ar[r]}
1705: &*+++[o][F]{{\bot}}
1706: \ar@/^/[ur]|{\textstyle{AB}}
1707: \ar[r]|{\textstyle{BC}}
1708: \ar@/_/[dr]|{\textstyle{AC}}
1709: &*+++[o][F]{{\bot}}
1710: \ar[ur]|(.3){\textstyle{H}}
1711: \ar[dr]|{\textstyle{T}}
1712: \\
1713: &&*+++[o][F]{{\bot}}
1714: \ar@/_/[r]_{\textstyle{H\lor T}}
1715: &*+++[o][F]{{\bot}}
1716: \ar@(ur,dr)[]
1717: }
1718: \]
1719: \caption[]{Moore machine of \eqref{3prisRR}. It is the minimalization
1720: of the Moore machine from Fig.\ \ref{f7}, so they are indeed logically
1721: indistinguishable.}
1722: \label{f3prisRR}
1723: \end{figure}
1724: 
1725: The general conclusion is that simple Moore machines can correspond to
1726: complicated $\TT$ formulas, and simple $\TT$ descriptions can yield
1727: complicated Moore machines. If we additionally take into account that
1728: it is hard to expect that any computer program will be ever able to
1729: transform human-readable representations of one kind into
1730: human-readable representations of the other kind\footnote{In both
1731: cases even graphical layout can have a huge impact on the readability
1732: of the model!}, we recommend that the whole process of modelling is
1733: done using only one of the formalisms, without mixing them.
1734:  
1735: \subsection{Algorithm for calculating the probability}\label{algorytmy}
1736: 
1737: 
1738: Of course, the natural method to compute probability of a given
1739: regular conditional $c$ in our model is to refer to an underlying
1740: Markov chain $\X,$ perform the computations there, and then use the
1741: formula
1742: 
1743: \[\Pr(c)=\frac{\sum_{i:h(i)=1}\lim_{n\to\infty}\Pr(X_n=i)}
1744: {\sum_{i:h(i)=1~\text{or}~0}\lim_{n\to\infty}\Pr(X_n=i)},
1745: \]
1746: 
1747: which follows directly from the Bayes' Formula.
1748: 
1749: The calculation of $\lim_{n\to\infty}\Pr(X_n=i)$ is generally known to
1750: be polynomial time in the number of states of the Markov chain,
1751: assuming unit cost of arithmetical operations \cite{KS}. The book
1752: \cite{stewart} contains the account of state-of-the-art algorithms
1753: for numerical calculations of the limiting probabilities.
1754: 
1755: As an example we calculate here the probability of the formula
1756: \eqref{3pris}, using the simplest possible approach, assuming that all
1757: the events from $\Omega$ have nonzero probability.
1758: 
1759: 
1760: We assume the following numbering of the states of the Markov chain
1761: from Fig.\ \ref{f9}:
1762: 
1763: 
1764: \begin{figure}[!hbtp]
1765: \[
1766: \xymatrix @R=0mm @C=0mm
1767: {
1768: &&*+++[o][F]{3}
1769: &&*+++[o][F]{5}
1770: \\
1771: &*+++[o][F]{1}
1772: &&&&*+++[o][F]{4}
1773: \\
1774: *+++[o][F]{2}
1775: &&&&&&*+++[o][F]{6}
1776: \\
1777: &*+++[o][F]{8}
1778: &&*+++[o][F]{7}
1779: &&*+++[o][F]{9}
1780: }
1781: \]
1782: \caption[]{Numbering of the states of Markov chain resulting from the
1783: Moore machine in Fig.\ \ref{f9}.}\label{f10}
1784: \end{figure}
1785: 
1786: 
1787: Then the matrix $\Pi$ of transition probabilities is
1788: \[
1789: \begin{bmatrix} 
1790: \bar{AB}& 0& 0& 
1791: \bar{BC}&0 & 0&
1792: \bar{AC}&0 & 0\\
1793: \bar{AB}& 0& 0& 
1794: \bar{BC}&0 & 0&
1795: \bar{AC}&0 & 0\\
1796: \bar{AB}& 0& 0& 
1797: \bar{BC}&0 & 0&
1798: \bar{AC}&0 & 0
1799:  \\ 
1800: 0& \bar{AB}\,\bar{H}& \bar{AB}\,\bar{T}&
1801: 0& \bar{BC}\,\bar{H}& \bar{BC}\,\bar{T}&
1802: 0& \bar{AC}\,\bar{H}& \bar{AC}\,\bar{T}\\
1803: 0& \bar{AB}\,\bar{H}& \bar{AB}\,\bar{T}&
1804: 0& \bar{BC}\,\bar{H}& \bar{BC}\,\bar{T}&
1805: 0& \bar{AC}\,\bar{H}& \bar{AC}\,\bar{T}\\
1806: 0& \bar{AB}\,\bar{H}& \bar{AB}\,\bar{T}&
1807: 0& \bar{BC}\,\bar{H}& \bar{BC}\,\bar{T}&
1808: 0& \bar{AC}\,\bar{H}& \bar{AC}\,\bar{T}
1809:  \\ 
1810: 0& 0& \bar{AB}&
1811: 0& 0& \bar{BC}&
1812: 0& 0& \bar{AC}\\
1813: 0& 0& \bar{AB}&
1814: 0& 0& \bar{BC}&
1815: 0& 0& \bar{AC}\\
1816: 0& 0& \bar{AB}&
1817: 0& 0& \bar{BC}&
1818: 0& 0& \bar{AC}
1819:       \end{bmatrix}
1820: \]
1821: where $\bar{AB}$ stands for $\pr(AB),$ and similarly for arguments
1822: $BC, AC, H,T$ (the matrix does not fit into the page when the standard
1823: notation is used).
1824: 
1825: It can be directly checked that the square of this matrix has all
1826: entries positive, hence the whole represents a single ergodic
1827: class. (This is what breaks down when some elements from $\Omega$ have
1828: probability $0.$ It this is permitted, one has to consider a few more
1829: cases.) It is known that in such cases the limiting probability does
1830: not depend on the initial probabilities of getting into this class,
1831: therefore we can ignore the dotted (transient) states from
1832: Fig.~\ref{f9}. The limiting probabilities can be found, given
1833: $\Pi=(p_{ij}),$ by finding the only solution of the system of linear
1834: equations
1835: 
1836: \[\left\{\begin{array}{lcl}
1837: \sum_{i=1}^9x_i&=&1,\\[5pt]
1838: \sum_{i=1}^9p_{i1}x_i&=&x_1,\\[5pt]
1839: \sum_{i=1}^9p_{i2}x_i&=&x_2,\\[5pt]
1840: \cdots&=&\cdots\\[5pt]
1841: \sum_{i=1}^9p_{i9}x_i&=&x_9,
1842:          \end{array}\right.
1843: \]
1844: 
1845: which yields the following unique solution:
1846: 
1847: \begin{align*}
1848: x_1&=\bar{AB}^{2}&
1849: x_2&=\bar{BC}\,\bar{AB}\,\bar{H}&
1850: x_3&= \bar{AB}(1-\bar{AB} - \bar{BC}\,\bar{H})\\
1851: x_4&=\bar{AB}\,\bar{BC}&
1852: x_5&=\bar{BC}^{2}\,\bar{H}&
1853: x_6&= \bar{BC}(1-\bar{AB} -\bar{BC}\,\bar{H})\\
1854: x_7&=\bar{AC}\,\bar{AB}&
1855: x_8&=\bar{BC}\,\bar{AC}\,\bar{H}&
1856: x_9&= 1 - \bar{AB}(1 + \bar{AC} + \bar{BC}\,\bar{H})- \bar{BC} 
1857: \end{align*}
1858: 
1859: and the asymptotic probability of the conditional represented by the
1860: Moore machine in question is $\dfrac{\Pr(AB)}{
1861: \Pr(BC)\Pr(H)+\Pr(AB)},$ as expected. In particular, in the
1862: equiprobable case the value is $2/3.$
1863: 
1864: \section{Related work and possible extensions}
1865: 
1866: 
1867: \subsection{Related work}\label{related}
1868: 
1869: \begin{itemize}
1870: \item
1871: Using temporal logic in reasoning about knowledge is nothing
1872: new. Indeed, many logics of knowledge incorporate temporal operators,
1873: see \cite{Fagin}. However, to the best of our knowledge, $\TT$ is the
1874: very first multi-valued temporal logic to be considered. In
1875: particular, the above mentioned logics of knowledge are two-valued.
1876: Moreover, $\TT$ is the first natural use of past tense temporal logic
1877: in computer science. Most of the established formalisms which use
1878: propositional temporal logic, indeed use its future tense fragment.
1879: 
1880: \item
1881: Computing of conditional probabilities $\Pr(\varphi|\psi)$ is not new,
1882: either, and has been considered by several authors, including
1883: \cite{liogonkii,halp1,halp2}, mostly for first order logic of
1884: unordered structures.
1885: 
1886: 
1887: \item
1888: Finally, Markov chains have already been used for evaluation of
1889: probabilities of logical statements. In particular, our Bayes' Formula
1890: is a simple extension of a theorem of Ehrenfeucht (see \cite{lynch}),
1891: phrased there as a theorem about first order logic of ordered unary
1892: structures (over which first order logic is equally as expressive as
1893: propositional temporal logic, see \cite{Emerson}).
1894: \end{itemize}
1895: 
1896: \subsection{Possible extensions.}
1897: 
1898: \begin{itemize}
1899: \item
1900: $\TT$ is not closed under its own connectives, since the nesting of
1901: the conditioning operator $(\cdot|\cdot)$ with other connectives (let
1902: alone itself) is not allowed, and since the temporal connectives
1903: cannot be applied to a conditional pair. As a consequence, operations
1904: on conditionals are defined by disassembling the pairs and
1905: reassembling them afterwards, to yield a pair in the correct
1906: syntactical form again.
1907: 
1908: We would like to have an equivalent logic with much better syntactical
1909: structure. This should be possible by extending the ideas of
1910: multivalued modal logics, investigated in \cite{mo,fit1,fit2}, by a
1911: multivalued counterparts of $\Since.$ The logic would then assume the
1912: form of a propositional logic with multivalued temporal connectives
1913: and conditioning.
1914: 
1915: The big question is whether one can retain the Bayes' Formula
1916: then. The existing attempts in the present tense logics of
1917: conditionals suggest it might be difficult.
1918: 
1919: \item 
1920: $\TT$ does not match exactly the class of automata, which for any
1921: assignment of probabilities yield a Markov chain with all states
1922: either transient or aperiodic. In such Markov chains all the limiting
1923: probabilities do exist, and thus every such Markov chain can be
1924: meaningfully considered to represent an extended kind of a
1925: conditional. Indeed, below is a simple example of such an automaton.
1926: 
1927: 
1928: \begin{figure}[!hbtp]
1929: \[\UseTips
1930: \xymatrix
1931: {
1932: *++[o][F]{~}
1933: \save[]+<-8mm,-4mm>*{}\ar[]\restore
1934: \ar@(l,u)^a
1935: \ar@/^/[r]^{a^\C}
1936: &*++[o][F]{~}
1937: \ar@(u,r)^{a^\C}
1938: \ar@/^/[d]^{a}
1939: \\
1940: *++[o][F]{~}
1941: \ar@(d,l)^{a^\C}
1942: \ar@/^/[u]^{a}
1943: &*++[o][F]{~}
1944: \ar@(r,d)^{a}
1945: \ar@/^/[l]^{a^\C}
1946: }
1947: \]
1948: \caption[]{It is not hard to verify
1949: that, no matter what probability is assigned to the event $a,$ the
1950: resulting Markov chain has only transient and acyclic states. However,
1951: the automaton is not acyclic, since it has two states, reachable by a
1952: path labelled $aa^\C$ from each other.}
1953: \end{figure}
1954: 
1955: 
1956: We would like to have an extension of $\TT,$ matching exactly the
1957: class of Markov chains with only transient and aperiodic states, to
1958: take the advantage of the maximal class of Markov chains for which the
1959: limiting probabilities exist, and thus all the definitions given in
1960: the paper make sense. We expect the logic to be obtained by extending
1961: the multivalued temporal logic proposed suggested above, rather than
1962: by extending the present syntax.
1963: 
1964: \end{itemize}
1965: 
1966: \paragraph{Acknowledgement.} The first author wishes to thank Igor
1967: Walukiewicz for valuable informations concerning temporal logic.
1968: 
1969: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1970: \newpage
1971: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1972: 
1973: \bibliographystyle{abbrv} \bibliography{bib}
1974: 
1975: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1976: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1977: 
1978: \end{document}
1979: % LocalWords:  cea
1980: 
1981: