1: \documentclass[amssymb,11pt]{article}
2: %\usepackage{hyperref} % this package must appear in first line
3: %\usepackage{latexsym} % for type $\Box$
4: %\usepackage{lgrind} % for formated source code
5: %\usepackage{latexcad} % latexcad.sty for drawings etc
6: \usepackage{graphicx}
7: %\usepackage{amsmath, amssymb, amsthm}
8: \setlength{\oddsidemargin}{10mm} \setlength{\marginparwidth}{0mm}
9: \setlength{\marginparsep}{0mm} \setlength{\topmargin}{0mm}
10: \setlength{\topskip}{0mm}
11:
12: \setlength{\textwidth}{140mm} \setlength{\textheight}{210mm}
13:
14: \setlength{\parskip}{2pt}
15:
16: \begin{document}
17: \title{A Theory of Computation Based on Quantum Logic (I)}
18: \author{Mingsheng Ying\thanks{This work was partly
19: supported by the National Key Project for Fundamental Research of
20: China (Grant No: 1998030905) and the National
21: Foundation of Natural Sciences of China (Grant No: 60273003)}\\
22: \\
23: \small \em State Key Laboratory of Intelligent Technology and Systems,\\
24: \small \em Department of Computer Science and Technology,\\
25: \small \em Tsinghua University, Beijing 100084, China,\\
26: \small \em Email: yingmsh@tsinghua.edu.cn}
27: \date{}
28: \maketitle
29:
30: \begin{abstract}
31: The (meta)logic underlying classical theory of computation is
32: Boolean (two-valued) logic. Quantum logic was proposed by Birkhoff
33: and von Neumann as a logic of quantum mechanics more than sixty
34: years ago. It is currently understood as a logic whose truth
35: values are taken from an orthomodular lattice. The major
36: difference between Boolean logic and quantum logic is that the
37: latter does not enjoy distributivity in general. The rapid
38: development of quantum computation in recent years stimulates us
39: to establish a theory of computation based on quantum logic. The
40: present paper is the first step toward such a new theory and it
41: focuses on the simplest models of computation, namely finite
42: automata. We introduce the notion of orthomodular lattice-valued
43: (quantum) automaton. Various properties of automata are carefully
44: reexamined in the framework of quantum logic by employing an
45: approach of semantic analysis. We define the class of regular
46: languages accepted by orthomodular lattice-valued automata. The
47: acceptance abilities of orthomodular lattice-valued
48: nondeterministic automata and their various modifications (such as
49: deterministic automata and automata with $\varepsilon-$moves) are
50: compared. The closure properties of orthomodular lattice-valued
51: regular languages are derived. The Kleene theorem about
52: equivalence of regular expressions and finite automata is
53: generalized into quantum logic. We also present a pumping lemma
54: for orthomodular lattice-valued regular languages. It is found
55: that the universal validity of many properties (for example, the
56: Kleene theorem, the equivalence of deterministic and
57: nondeterministic automata) of automata depend heavily upon the
58: distributivity of the underlying logic. This indicates that these
59: properties does not universally hold in the realm of quantum
60: logic. On the other hand, we show that a local validity of them
61: can be recovered by imposing a certain commutativity to the
62: (atomic) statements about the automata under consideration. This
63: reveals an essential difference between the classical theory of
64: computation and the computation theory based on quantum logic.
65: \end{abstract}\par
66:
67: \smallskip\
68:
69: \emph{Key Words:} Quantum logic, orthomodular lattice, algebraic
70: semantics,
71:
72: finite automata, regular languages, quantum computation
73:
74: \vspace{2em}
75:
76: \textbf{Contents}
77:
78: \smallskip\
79:
80: 1. Introduction (page 2)
81:
82: 2. Quantum logic (page 11)
83:
84: \ \ \ \ 2.1. Orthomodular lattices (page 11)
85:
86: \ \ \ \ 2.2. The language of quantum logic (page 20)
87:
88: \ \ \ \ 2.3. The algebraic semantics of quantum logic (page 20)
89:
90: \ \ \ \ 2.4. The operations of quantum sets (page 22)
91:
92: 3. Orthomodular lattice-valued automata (page 25)
93:
94: 4. Orthomodular lattice-valued deterministic automata (page 33)
95:
96: 5. Orthomodular lattice-valued automata with $\epsilon-$moves
97: (page 40)
98:
99: 6. Closure properties of orthomodular lattice-valued regularity
100: (page 44)
101:
102: 7. Orthomodular lattice-valued regular expressions (page 57)
103:
104: 8. Pumping lemma for orthomodular lattice-valued regular languages
105: (page 69)
106:
107: 9. Conclusion (page 71)
108:
109: \ \ \ \ References (page 73)
110:
111: \bigskip\
112:
113: \smallskip\
114:
115: \textbf{1. Introduction}
116:
117: \smallskip\
118:
119: It is well-known that an axiomatization of a mathematical theory
120: consists of a system of fundamental notions as well as a set of
121: axioms about these notions. The mathematical theory is then the
122: set of theorems which can be derived from the axioms. Obviously,
123: one needs a certain logic to provide tools for reasoning in the
124: derivation of these theorems from the axioms. As pointed out by A.
125: Heyting [He63, page 5], in elementary axiomatics logic was used in
126: an unanalyzed form. Afterwards, in the studies for foundations of
127: mathematics beginning in the early of twentieth century, it had
128: been realized that a major part of mathematics has to exploit the
129: full power of classical (Boolean) logic [Ha82], the strongest one
130: in the family of existing logics. For example, group theory is
131: based on first-order logic, and point-set topology is built on a
132: fragment of second-order logic. However, a few mathematicians,
133: including the big names L. E. J. Brouwer, H. Poincare, L.
134: Kronecker and H. Weyl, took some kind of constructive position
135: which is in more or less explicit opposition to certain forms of
136: mathematical reasoning used by the majority of the mathematical
137: community. Some of them even endeavored to establish so-called
138: constructive mathematics, the part of mathematics that could be
139: rebuilt on constructivist principles. The logic employed in the
140: development of constructive mathematics is intuitionistic logic
141: [TD88] which is truly weaker than classical logic.
142:
143: Since many logics different from classical logic and
144: intuitionistic logic have been invented in the last century, one
145: may naturally ask the question whether we are able to establish
146: some mathematical theories based on other nonclassical logics
147: besides intuitionistic logic. Indeed, as early as the first
148: nonclassical logics appeared, the possibility of building
149: mathematics upon them was conceived. As mentioned by A. Mostowski
150: [M65], J. Lukasiewicz hoped that there would be some nonclassical
151: logics which can be properly used in mathematics as non-Euclidean
152: geometry does. In 1952, J. B. Rosser and A. R. Turquette [RT52,
153: page 109] proposed a similar and even more explicit idea:
154: $$$$
155: \textit{"The fact that it is thus possible to generalize the
156: ordinary two-valued logic so as not only to cover the case of
157: many-valued statement calculi, but of many-valued quantification
158: theory as well, naturally suggests the possibility of further
159: extending our treatment of many-valued logic to cover the case of
160: many-valued sets, equality, numbers, etc. Since we now have a
161: general theory of many-valued predicate calculi, there is little
162: doubt about the possibility of successfully developing such
163: extended many-valued theories. ... we shall consider their careful
164: study one of the major unsolved problems of many-valued logic."}
165: $$$$
166: Unfortunately, the above idea has not attracted much attention in
167: logical community. For such a situation, A. Mostowski [M65]
168: pointed out that most of nonclassical logics invented so far have
169: not been really used in mathematics, and intuitionistic logic
170: seems the unique one of nonclassical logics which still has an
171: opportunity to carry out the Lukasiewicz's project. A similar
172: opinion was also expressed by J. Dieudonne [Di78], and he said
173: that mathematical logicians have been developing a variety of
174: nonclassical logics such as second-order logic, modal logic and
175: many-valued logic, but these logics are completely useless for
176: mathematicians working in other research areas.
177:
178: One reason for this situation might be that there is no suitable
179: method to develop mathematics within the framework of nonclassical
180: logics. As was pointed out above, classical logic is applied as
181: the deduction tool in almost all mathematical theories. It should
182: be noted that what is used in these theories is the deductive
183: (proof-theoretical) aspect of classical logic. However, the proof
184: theory of nonclassical logics is much more complicated than that
185: of classical logic, and it is not an easy task to conduct
186: reasoning in the realm of the proof theory of nonclassical logics.
187: It is the case even for the simplest nonclassical logics,
188: three-valued logics. This is explicitly indicated by the following
189: excerpt from H. Hodes [Ho89]:
190: $$$$
191: \textit{"Of course three-valued logics will be somewhat more
192: complicated than classical two-valued logic. In fact,
193: proof-theoretically they are at least twice as complicated: ....
194: But model-theoretically they are only 50 percent more
195: complicated,....}
196: $$$$
197: And much worse, some nonclassical logics were introduced only in a
198: semantic way, and the axiomatizations of some among them are still
199: to be found, and some of them may be not (finitely) axiomatizable.
200: Thus, our experience in studying classical mathematics may be not
201: suited, or at least cannot directly apply, to develop mathematics
202: based on nonclassical logics. In the early 1990's an attempt had
203: been made by the author [Yi91-93; Yi93] to give a partial and
204: elementary answer in the case of point-set topology to the J. B.
205: Rosser and A. R. Turquette's question raised above. We employed a
206: semantical analysis approach to establish topology based on
207: residuated lattice-valued logic, especially the Lukasiewicz system
208: of continuous-valued logic. Roughly speaking, the semantical
209: analysis approach transforms our intended conclusions in
210: mathematics, which are usually expressed as implication formulas
211: in our logical language, into certain inequalities in the
212: truth-value lattice by truth valuation rules, and then we
213: demonstrate these inequalities in an algebraic way and conclude
214: that the original conclusions are semantically valid. We believe
215: that semantical analysis approach is an effective method to
216: develop mathematics based on nonclassical logics.
217:
218: A much more essential reason for the situation that few
219: nonclassical logics have been applied in mathematics is absence of
220: appealing from other subjects or applications in the real world.
221: One major exception may be the case of quantum logic. Quantum
222: logic was introduced by G. Birkhoff and J. von Neumann [BN36] in
223: the thirties of the twentieth century as the logic of quantum
224: mechanics. They realized that quantum mechanical systems are not
225: governed by classical logical laws. Their proposed logic stems
226: from von Neumann's Hilbert space formalism of quantum mechanics.
227: The starting point was explained very well by the following
228: excerpt from G. Birkhoff and J. von Neumann [BN36]:
229: $$$$
230: \textit{"what logical structure one may hope to find in physical
231: theories which, like quantum mechanics, do not conform to
232: classical logic. Our main conclusion, based on admittedly
233: heuristic arguments, is that one can reasonably expect to find a
234: calculus of propositions which is formally indistinguishable from
235: the calculus of linear subspaces [of Hilbert space] with respect
236: to set products, linear sums, and orthogonal complements - and
237: resembles the usual calculus of propositions with respect to
238: 'and', 'or', and 'not'."}
239: $$$$
240: Thus linear (closed) subspaces of Hilbert space are identified
241: with propositions concerning a quantum mechanical system, and the
242: operations of set product, linear sum and orthogonal complement
243: are treated as connectives. By observing that the set of linear
244: subspaces of a finite-dimensional Hilbert space together with
245: these operations enjoys Dedekind's modular law, G. Birkhoff and J.
246: von Neumann [BN36] suggested to use modular lattices as the
247: algebraic version of the logic of quantum mechanics, just like
248: that Boolean algebras act as an algebraic counterpart of classical
249: logic. However, the modular law does not hold in an
250: infinite-dimensional Hilbert space. In 1937, K. Husimi [Hu37]
251: found a new law, called now the orthomodular law, which is valid
252: for the set of linear subspaces of any Hilbert space. Nowadays,
253: what is usually called quantum logic in the mathematical physics
254: literatures refers to the theory of orthomodular lattices.
255: Obviously, this kind of quantum logic is not very logical. Indeed,
256: there is also another much more 'logical' point of view on quantum
257: logic in which quantum logic is seen as a logic whose truth values
258: range over an orthomodular lattice (for an excellent exposition
259: for the latter approach of quantum logic, see M. L. Dalla Chiara
260: [DC86], or J. P. Rawling and S. A. Selesnick [RS00]). After the
261: invention of quantum logic, quite a few mathematicians have tried
262: to establish mathematics based on quantum logic. Indeed, J. von
263: Neumann [N62] himself proposed the idea of considering a quantum
264: set theory, corresponding to quantum logic, as does classical set
265: theory to classical logic. One important contribution in this
266: direction was made by G. Takeuti [T81]. His main idea was
267: explained, and the nature of mathematics based on quantum logic
268: was analyzed very well by the following citation from the
269: introduction of [T81]:
270: $$$$
271: \textit{"Since quantum logic is an intrinsic logic, i.e. the logic
272: of the quantum world, it is an important problem to develop
273: mathematics based on quantum logic, more specifically set theory
274: based on quantum logic. It is also a challenging problem for
275: logicians since quantum logic is drastically different from the
276: classical logic or the intuitionistic logic and consequently
277: mathematics based on quantum logic is extremely difficult. On the
278: other hand, mathematics based on quantum logic has a very rich
279: mathematical content. This is clearly shown by the fact that there
280: are many complete Boolean algebras inside quantum logic. For each
281: complete Boolean algebra $B$, mathematics based on $B$ has been
282: shown by our work on Boolean valued analysis to have rich
283: mathematical meaning. Since mathematics based on $B$ can be
284: considered as a sub-theory of mathematics based on quantum logic,
285: there is no doubt about the fact that mathematics based on quantum
286: logic is very rich. The situation seems to be the following.
287: Mathematics based on quantum logic is too gigantic to see through
288: clearly."}
289: $$$$
290: The main technical result of G. Takeuti [T81] is a construction of
291: orthomodular lattice-valued universe. He built up such an universe
292: in a way similar to Boolean-valued models of ZF + AC, and showed
293: that a reasonable set theory, including some axioms from ZF + AC
294: or their slight modifications, holds in this universe. Recently,
295: K. -G. Schlesinger [Sc99] developed a theory of quantum sets by
296: using a categorical approach in the spirit of topos theory. He
297: started with the category of complex (pre-)Hilbert spaces and
298: linear maps. This category was seen as the (basic) quantum set
299: universe. Then he was able to introduce the analog of number
300: systems and to deal with the analog of some algebraic structures
301: in quantum set theory. Indeed, K. -G. Schlesinger's terminal goal
302: is to build a quantum mathematics, i.e., a mathematical theory
303: where all the ingredients (like logic and set theory) adhere to
304: the rules of quantum mechanics. Quantum set theory is the
305: quantization of the mathematical theory of pure objects, and so it
306: is just the first step toward his goal. It is worth noting that
307: the role of quantum logic in such a quantum mathematics is
308: different from that in G. Takeuti's quantum set theory, and
309: quantum logic appears as an internal logic in the former.
310:
311: After a careful examination on the development of mathematics
312: based on nonclassical logics, we now come to explore the
313: possibility of establishing a theory of computation based on
314: nonclassical logics. A formal formulation of the notion of
315: computation is one of the greatest scientific achievements in the
316: twentieth century. Since the middle of 1930's, various models of
317: computation have been introduced, such as Turing machines, Post
318: systems, $\lambda-$calculus and $\mu-$recursive functions. In
319: classical computing theory, these models of computations are
320: investigated in the framework of classical logic; more explicitly,
321: all properties of them are deduced by classical logic as a
322: (meta)logical tool. So, it is reasonable to say that classical
323: computing theory is a part of classical mathematics. Knowing the
324: basic idea of mathematics based on nonclassical logics, we may
325: naturally ask the question: is it possible to build a theory of
326: computation based on nonclassical logics, and what are the same of
327: and difference between the properties of the models of
328: computations in classical logic and the corresponding ones in
329: non-classical logics? There has been a very big population of
330: non-classical logics. Of course, it is unnecessary to construct
331: models of computations in each nonclassical logic and to compare
332: them with the ones in classical logic because some nonclassical
333: logics are completely irrelative to behaviors of computations.
334: Nevertheless, as will be explained shortly, it is absolutely worth
335: studying deeply and systematically models of computations based on
336: quantum logic.
337:
338: It seems that both points of views on quantum logic mentioned
339: above have no obvious links to computations; but appearance of the
340: idea of quantum computers changed dramatically the long-standing
341: situation. The idea of quantum computation came from the studies
342: of connections between physics and computation. The first step
343: toward it was the understanding of the thermodynamics of classical
344: computation. In 1973, C. H. Bennet [Ben73] noted that a logically
345: reversible operation need not dissipate any energy and found that
346: a logically reversible Turing machine is a theoretical
347: possibility. In 1980, further progress was made by P. A. Benioff
348: [Be80] who constructed a quantum mechanical model of Turing
349: machine. His construction is the first quantum mechanical
350: description of computer, but it is not a real quantum computer. It
351: should be noted that in P. A. Benioff's model between computation
352: steps the machine may exist in an intrinsically quantum state, but
353: at the end of each computation step the tape of the machine always
354: goes back to one of its classical states. Quantum computers were
355: first envisaged by R. P. Feynman [Fe82; Fe86]. In 1982, he [Fe82]
356: conceived that no classical Turing machine could simulate certain
357: quantum phenomena without an exponential slowdown, and so he
358: realized that quantum mechanical effects should offer something
359: genuinely new to computation. Although R. P. Feynman proposed the
360: idea of universal quantum simulator, he did not give a concrete
361: design of such a simulator. His ideas were elaborated and
362: formalized by D. Deutsch in a seminal paper [De85]. In 1985, D.
363: Deutsch described the first true quantum Turing machine. In his
364: machine, the tape is able to exist in quantum states too. This is
365: different from P. A. Benioff's machine. In particular, D. Deutsch
366: introduced the technique of quantum parallelism by which quantum
367: Turing machine can encode many inputs on the same tape and perform
368: a calculation on all the inputs simultaneously. Furthermore, he
369: proposed that quantum computers might be able to perform certain
370: types of computations that classical computers can only perform
371: very inefficiently. One of the most striking advances was made by
372: P. W. Shor [S94] in 1994. By exploring the power of quantum
373: parallelism, he discovered a polynomial-time algorithm on quantum
374: computers for prime factorization of which the best known
375: algorithm on classical computers is exponential. In 1996, L. K.
376: Grover [Gr96] offered another apt killer of quantum computation,
377: and he found a quantum algorithm for searching a single item in an
378: unsorted database in square root of the time it would take on a
379: classical computer. Since both prime factorization and database
380: search are central problems in computer science and the quantum
381: algorithms for them are highly faster than the classical ones, P.
382: W. Shor and L. K. Grover's works stimulated an intensive
383: investigation on quantum computation. After that, quantum
384: computation has been an extremely exciting and rapidly growing
385: field of research.
386:
387: The studies of quantum computation may be roughly divided into
388: four strata, arranged according increasing order of abstraction
389: degree: (1) physical implementations; (2) physical models; (3)
390: mathematical models; and (4) logical foundations. Almost all
391: pioneer works such as [Be80, F82, D85] in this field were devoted
392: to build physical models of quantum computing. In 1990's, a great
393: attention was paid to the physical implementation of quantum
394: computation. For example, S. Lloyd [L93] considered the practical
395: implementation by using electromagnetic pulses and J. I. Cirac and
396: P. Zoller [CZ95] used laser manipulations of cold trapped ions to
397: implement quantum computing. The current theoretical concerns in
398: the area of quantum computation have mainly been given to quantum
399: algorithms. But also there have been a few attempts to develop
400: mathematical models of quantum computation and to clarify the
401: relationship between different models. For example, except quantum
402: Turing machines, D. Deutsch [De89] also proposed the quantum
403: circuit model of computation, and A. C. Yao [Ya93] showed that the
404: quantum circuit model is equivalent to the quantum Turing machine
405: in the sense that they can simulate each other in polynomial time.
406: As is well known, in classical computing theory, there are still
407: two important classes of models of computation rather than Turing
408: machines; namely, finite automata and pushdown automata. They are
409: equipped with finite memory or finite memory with stack,
410: respectively, and so have weaker computing power than Turing
411: machines. Recently, J. P. Crutchfield and C. Moore [CM00], A.
412: Kondacs and J. Watrous [KW97], and S. Gudder [Gu00] tried to
413: introduce some quantum devices corresponding to these weaker
414: models of computation. Roughly speaking, quantum automata may be
415: seen as quantum counterparts of probabilistic automata. In a
416: probabilistic automaton, each transition is equipped with a number
417: in the unit interval to indicate the probability of the occurrence
418: of the transition; by contrast in a quantum automaton we associate
419: with each transition a vector in a Hilbert space which is
420: interpreted as the probability amplitude of the transition. In a
421: sense, these mathematical models of quantum computation can be
422: seen as abstractions of its physical models.
423:
424: It should be noted that the theoretical models of quantum
425: computation mentioned above, including quantum Turing machines and
426: quantum automata, are still developed in classical (Boolean)
427: logic. Thus, their logical basis is the same as that of classical
428: computation, and we may argue that sometimes these models might be
429: not suitable for quantum computers that obey some logical laws
430: different from that in Boolean logic. Indeed, V. Vedral and M. B.
431: Plenio [VP98] already advocated that quantum computers require
432: quantum logic, something fundamentally different to classical
433: Boolean logic. As stated above, quantum logic has been existing
434: for a long time. So, the point is how to apply quantum logic in
435: the analysis and design of quantum computers. The background
436: exposed above highly motivates us to explore the possibility of
437: establishing a theory of computation based on quantum logic. The
438: purpose of the present paper and its continuations is exactly to
439: develop such a new theory. In a sense, our approach may be thought
440: of as a logical foundation of quantum computation and a further
441: abstraction of its mathematical models. The relation between
442: Crutchfield et al's studies [CM00, KW97, Gu00] on quantum automata
443: and our automata theory based on quantum logic is quite similar to
444: that between J. von Neumann's Hilbert space formalism of quantum
445: mechanics and quantum logic.
446:
447: Since finite automata are the simplest models of computation (with
448: finite memory), in this paper we focus our attention on developing
449: a theory of finite automata based on quantum logic. The present
450: paper is organized as follows. In Section 2, we recall some basic
451: notions and results of quantum logic and its algebraic semantics
452: needed in the subsequent sections from the previous literature.
453: Some new lemmas on implication operators in quantum logic are
454: presented too. They are crucial in the proofs of several main
455: results in this paper. In Section 3, we introduce the notion of
456: orthomodular lattice-valued (quantum) automaton. Then two
457: different orthomodular lattice-valued predicates of regularity on
458: languages are proposed. These two predicates stands indeed for the
459: (orthomodular lattice-valued) class of languages accepted by
460: orthomodular lattice-valued automata. This provides us with a
461: framework in which various properties of automata can be
462: reexamined within quantum logic. The technique employed in this
463: paper is mainly the approach of semantic analysis developed in
464: [Yi91-93; Yi93]. The acceptance ability of orthomodular
465: lattice-valued nondeterministic automata are then compared with
466: that of their two kinds of modifications, namely deterministic
467: automata and automata with $\varepsilon-$moves, respectively in
468: Sections 4 and 5. The closure properties of orthomodular
469: lattice-valued regular languages are derived in Section 6. In
470: Section 7, we introduce the notion of orthomodular lattice-valued
471: regular expression, and the Kleene theorem about equivalence of
472: regular expressions and finite automata is generalized into
473: quantum logic. Section 8 is devoted to present a pumping lemma for
474: orthomodular lattice-valued regular languages. Some basic ideas of
475: this paper were announced in [Yi00], and Definitions 3.1 and 3.2,
476: Examples 3.1-4, and Propositions 6.1 and 6.3 were also presented
477: there. For completeness, however, they are included in the present
478: paper.
479:
480: The most interesting thing is, in the author's opinion, the
481: discovery that the universal validity of many properties (for
482: example, the Kleene theorem, the equivalence of deterministic and
483: nondeterministic automata) of automata depend heavily upon the
484: distributivity of the underlying logic. It is shown that the
485: universal validity of these properties is equivalent to the
486: requirement that the set of truth values of the meta-logic
487: underlying our theory of automata is a Boolean algebra. This
488: indicates that these properties does not universally hold in the
489: realm of quantum logic, and it is in fact a negative conclusion in
490: our theory of automata based on quantum logic. Furthermore, it
491: implies the fact that an essential difference exists between the
492: classical theory of computation and the computation theory based
493: on quantum logic.
494:
495: Observing that some important properties of automata cannot be
496: built within quantum logic, one may naturally ask the question
497: whether they may be partially recast without appealing to
498: distributivity of the underlying logic. Fortunately, we are able
499: to show that a local validity of these properties of automata can
500: be recovered by imposing a certain commutativity to the truth
501: values of the (atomic) statements about the automata under
502: consideration. Very surprisingly, almost all results in classical
503: automata theory that are not valid in a non-distributive logic can
504: be revived by a certain commutativity in quantum logic. This
505: further leads us to a new question: why commutativity plays such a
506: key role for quantum automata, and is there any physical
507: interpretation for it? To answer this question, let us first note
508: that all truth values in quantum logic are taken from an
509: orthomodular lattice. The prototype of orthomodular lattice is the
510: set of linear (closed) subspaces of a Hilbert space with the set
511: inclusion as its ordering relation. Suppose that $X$ and $Y$ are
512: two subspaces of a Hilbert space $H$. Moreover, we use $P_X$ and
513: $P_Y$ to denote the projections on $X$ and $Y$ respectively. Then
514: $P_X$ and $P_Y$ are Hermitian operators on $H$, and they may be
515: seen as two (physical) observables $A$ and $B$ in a quantum system
516: whose state space is $H$, according to the Hilbert space formalism
517: of quantum mechanics. If we write $\Delta (A)$ and $\Delta (B)$
518: for the respective standard deviations of measurement on $A$ and
519: $B$, then the Heisenberg uncertainty principle gives the following
520: inequality:
521: $$\Delta (A)\cdot \Delta (B)\geq \frac{1}{2} |<\psi
522: |[A,B]|\psi>|$$ for all quantum state $|\psi>$ in $H$, where
523: $[A,b]=AB-BA$ is the commutator between $A$ and $B$. We now turn
524: back to the orthomodular lattice of the linear subspaces of $H$.
525: The commutativity of $A$ and $B$ is defined by the condition
526: $X=(X\wedge Y)\vee (X\wedge Y^{\bot})$, where $\wedge$, $\vee$ and
527: $\bot$ are respectively the meet, union and orthocomplement. It
528: may be seen that the commutativity between $X$ and $Y$ is
529: equivalent to exactly the fact that $A$ and $B$ commutate, i.e.,
530: $AB=BA$. In this case, $|<\psi|[A,B]|\psi>|=0$, and $\Delta
531: (A)\cdot \Delta (B)$ may vanish; or in other words, $\Delta (A)$
532: and $\Delta (B)$ can simultaneously become arbitrarily small.
533: Remember that in our theory of automata based on quantum logic the
534: commutativity is attached to the basic statements describing the
535: considered automata. On the other hand, the basic statements are
536: indeed corresponding to some actions in these automata. Therefore,
537: a potential physical interpretation for the need of commutativity
538: is that some nice properties of automata require the standard
539: deviations of the observables concerning the basic actions in
540: these automata being able to reach simultaneously very small
541: values.
542:
543: The results gained in our approach may offer some new insights on
544: the theory of computation. As an example, let us consider the
545: Church-Turing thesis. The realization that the intuitive notion of
546: "effective computation" can be identified with the mathematical
547: concept of "computation by the Turing machine" is based on the
548: fact that the Turing machine is computationally equivalent to some
549: vastly dissimilar formalisms for the same purpose, such as Post
550: systems, $\mu-$recursive functions, $\lambda-$calculus and
551: combinatory logic. As pointed out by J. E. Hopcroft and J. D.
552: Ullman [HU79], another reason for the acceptance of the Turing
553: machine as a general model of a computation is that the Turing
554: machine is equivalent to its many modified versions that would
555: seem off-hand to have increased computing power. We should note
556: that the equivalence between the Turing machine and its various
557: generalizations as well as other formalisms of computation has
558: been reached in classical Boolean logic. In addition, quantum
559: logic is known to be strictly weaker than Boolean logic. Thus, it
560: is reasonable to doubt that the same equivalence can be achieved
561: when our underlying meta-logic is replaced by quantum logic, and
562: the Church-Turing thesis needs to be reexamined in the realm of
563: quantum logic. Indeed, in a continuation of this paper we are
564: going to establish a theory of Turing machines based on quantum
565: logic. The details of such a theory is still to be exploited, but
566: the conclusion concerning the equivalence between deterministic
567: and nondeterministic automata obtained in this paper suggests us
568: to believe that the equivalence between deterministic and
569: nondeterministic Turing machines also depends upon the
570: distributivity of the underlying logic, and a certainty
571: commutativity for the basic actions in Turing machines will
572: guarantee such an equivalence. Keeping this belief in mind, we may
573: assert that a certain commutativity of the observables for some
574: basic actions in the Turing machine is a physical support of the
575: Church-Turing thesis in the framework of quantum logic.
576: Furthermore, with the above physical interpretation for
577: commutativity, this hints that there might be a deep connection
578: between the Heisenberg uncertainty principle and the Church-Turing
579: thesis, two of the greatest scientific discoveries in the
580: twentieth century. It is notable that such a connection could be
581: observed via an argument in a nonclassical logic (and it is
582: impossible to be found if we always work within the classical
583: logic). As early as in 1985, it was argued by D. Deutsch [De84]
584: that underlying the Church-Turing thesis there is an implicit
585: physical assertion. There is certainly no doubt about the
586: existence of such a physical assertion. The true problem here is:
587: what is it? The answer given by D. Deutsch himself is the
588: following physical principle: "every finitely realizable physical
589: system can be perfectly simulated by a universal model computing
590: machine operating by finite means". Our above analysis on the role
591: of commutativity in computation theory based on quantum logic
592: perhaps indicates that in order to be simulated by a universal
593: computing machine some observables of the physical system are
594: required to possess a certain commutativity. So, it is fair to say
595: that the observation on commutativity presented above provides a
596: complement to D. Deutsch's argument from a logical point of view.
597:
598: \bigskip\
599:
600: \textbf{2. Quantum Logic}
601:
602: \smallskip\
603:
604: The aim of this section is to recall some basic notions and
605: results about quantum logic needed in the subsequent sections and
606: to fix notations. In this paper, quantum logic is understood as a
607: complete orthomodular lattice-valued logic. This section is mainly
608: concerned with the semantic aspect of such a logic, and it will be
609: divided into four subsections. The first subsection will briefly
610: review some fundamental results on orthomodular lattices; for more
611: details, we refer to [Ka83] and [BH00]. In the second one we will
612: introduce the language of first-order quantum logic. The third
613: will discuss the algebraic semantics of first-order quantum logic.
614: Some useful properties of orthomodular lattice-valued sets are
615: given in the fourth subsection.
616:
617: \smallskip\
618:
619: \textbf{2.1. Orthomodular Lattices}
620:
621: \smallskip\
622:
623: The set of truth values of a quantum logic will be taken to be an
624: orthomodular lattice. So we first introduce the notion of
625: orthomodular lattice. An ortholattice is a 7-tuple
626: $$\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$$ where:
627:
628: (1) $<L,\leq ,\wedge ,\vee ,0,1>$ is a bounded lattice, $0,1$ are
629: the least and greatest elements of $L,$ respectively, $\leq $ is
630: the partial ordering in $L,$ and for any $a,b\in L,$ $a\wedge b,$
631: and $a\vee b$ stand for the greatest lower bound and the least
632: upper bound of $a$ and $b$, respectively;
633:
634: (2) $\bot $ is a unary operation on $L,$ called orthocomplement,
635: and required to satisfy the following conditions: for any $a,b\in
636: L$,
637:
638: \ \ \ (i) $a\wedge a^{\bot }=0,\ a\vee a^{\bot }=1;$
639:
640: \ \ \ (ii) $a^{\bot \bot }=a;$ and
641:
642: \ \ \ (iii) $a\leq b$ implies $b^{\bot }\leq a^{\bot }.$
643:
644: It is easy to see that the condition (iii) is equivalent to one of
645: the De Morgan laws: for any $a,b\in L$,
646:
647: \ \ \ (iii') $(a\wedge b)^{\bot} = a^{\bot}\vee b^{\bot},\ (a\vee
648: b)^{\bot} = a^{\bot}\wedge b^{\bot}.$
649:
650: Let $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an ortholattice,
651: and let $a,b\in L$. We say that $a$ commutes with $b$, in symbols
652: $aCb$, if $$a=(a\wedge b)\vee (a\wedge b^{\bot}).$$
653:
654: An orthomodular lattice is an ortholattice $\ell =<L,\leq ,\wedge
655: ,\vee ,\bot ,0,1>$ satisfying the orthomodular law: for all
656: $a,b\in L$,
657:
658: \ \ \ (iv) $a\leq b\ {\rm implies}\ a\vee (a^{\bot}\wedge b)=b.$
659:
660: The orthomodular law can be replaced by the following equation:
661:
662: \ \ \ (iv') $a\vee (a^{\bot }\wedge (a\vee b))=a\vee b\ {\rm for\
663: any}\ a,b\in L.$
664:
665: A Boolean algebra is an ortholattice $\ell =<L,\leq ,\wedge ,\vee
666: ,\bot ,0,1>$ fulfilling the distributive law of join over meet:
667: for all $a,b,c\in L$,
668:
669: \ \ \ (v) $a\vee (b\wedge c)=(a\vee b)\wedge (a\vee c).$
670:
671: With the De Morgan law it is easy to know that the condition (v)
672: is equivalent to the distributive law of meet over join: for any
673: $a,b,c\in L$,
674:
675: \ \ \ (v') $a\wedge (b\vee c)=(a\wedge b)\vee (a\wedge c).$
676:
677: Obviously, the distributive law implies the orthomodular law, and
678: so a Boolean algebra is an orthomodular lattice.
679:
680: The following lemma gives a characterization of orthomodular
681: lattices and it distinguishes orthomodular lattices from
682: ortholattices.
683:
684: \smallskip\
685:
686: \textbf{Lemma 2.1.} ([BH00], Propositions 2.1 and 2.2) Let $\ell
687: =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an ortholattice. Then the
688: following seven statements are equivalent:
689:
690: (1) $\ell$ is an orthomodular lattice;
691:
692: (2) For any $a,b\in L$, if $a\leq b$ and $a^{\bot}\wedge b=0$ then
693: $a=b$;
694:
695: (3) For any $a,b\in L$, if $aCb$ then $bCa$;
696:
697: (4) For any $a,b\in L$, if $aCb$ then $a^{\bot}Cb$;
698:
699: (5) For any $a,b\in L$, if $aCb$ then $a\vee (a^{\bot}\wedge
700: b)=a\vee b$;
701:
702: (6) The benzene ring $O_6$ (see Figure 1) is not a subalgebra of
703: $\ell$.
704: \begin{figure}\centering
705: \includegraphics{fig_1.eps}
706: \caption{Benzene ring} \label{fig 1 }
707: \end{figure}
708:
709: (7) For any $a,b\in L$, if $a\leq b$ then the subalgebra $[a,b]$
710: of $\ell$ generated by $a$ and $b$ is a Boolean
711: algebra.$\heartsuit$
712:
713: \smallskip\
714:
715: The set of truth values of classical logic is a Boolean algebra;
716: whereas quantum logic is an orthomodular lattice-valued logic. It
717: is well-known that a Boolean algebra must be an orthomodular
718: lattice, but the inverse is not true. Thus, quantum logic is
719: weaker than classical logic. The major difference between a
720: Boolean algebra and an orthomodular lattice is that distributivity
721: is not valid in the latter. However, many cases still appeal an
722: application of the distributivity even when we manipulate elements
723: in an orthomodular lattice. This requires us to regain a certain
724: (weaker) version of distributivity in the realm of orthomodular
725: lattices. The key technique for this purpose is commutativity
726: which is able to provide a localization of distributivity. The
727: following lemma together with Lemma 2.1(4) indicates that
728: commutativity is preserved by all operations of orthomodular
729: lattice.
730:
731: \smallskip\
732:
733: \textbf{Lemma 2.2}. ([BH00], Proposition 2.4) Let $\ell =<L,\leq
734: ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular lattice, and let
735: $a\in L$ and $b_i\in L$ $(i\in I)$. If $aCb_i$ for any $i\in I$,
736: then $$aC\wedge_{i\in I}b_i\ {\rm and}\ aC\vee_{i\in I}b_i$$
737: provided $\wedge_{i\in I}b_i$ and $\vee_{i\in I}b_i$
738: exist.$\heartsuit$
739:
740: \smallskip\
741:
742: The local distributivity implied by commutativity is then given by
743: the following
744:
745: \smallskip\
746:
747: \textbf{Lemma 2.3}. ([BH00], Proposition 2.3) Let $\ell =<L,\leq
748: ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular lattice and let
749: $A\subseteq L$. For any $a\in A$ and $b_i\in A$ $(i\in I)$, if
750: $aCb_i$ for all $i\in I$, then
751: $$a\wedge \vee_{i\in I}b_i=\vee_{i\in I}(a\wedge b_i),$$
752: $$a\vee \wedge_{i\in I}b_i=\wedge_{i\in I}(a\vee b_i)$$
753: provided $\wedge_{i\in I} b_i$ and $\vee_{i\in I} b_i$
754: exist.$\heartsuit$
755:
756: \smallskip\
757:
758: The above lemma is very useful, and it often enables us to recover
759: distributivity in an orthomodular lattice. However, its condition
760: that all elements involved commute each other is quite strong, and
761: not easy to meet. This suggests us to find a way to weaken this
762: condition. One solution was found by G. Takeuti [T81], and he
763: introduced the notion of commutator which can be seen as an index
764: measuring the degree to which the commutativity is valid.
765:
766: \smallskip\
767:
768: \textbf{Definition 2.1}. ([T81], pages 305 and 307) Let $\ell
769: =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular lattice and
770: let $A\subseteq L$.
771:
772: (1) If $A$ is finite, then the commutator $\gamma (A)$ of $A$ is
773: defined by
774: $$\gamma (A)=\vee\{\wedge_{a\in A}a^{f(a)}:f\ {\rm is\ a\ mapping\ from}\ A\ {\rm into}\
775: \{1,-1\}\},$$ where $a^{1}$ denotes $a$ itself and $a^{-1}$
776: denotes $a^{\bot}$.
777:
778: (2) The strong commutator $\Gamma (A)$ of $A$ is defined by
779: $$\Gamma (A)=\vee\{b: C(a,b)\ {\rm for\ all}\ a\in A\ {\rm and}\
780: C(a_1\wedge b, a_2\wedge b)\ {\rm for\ all}\ a_1, a_2\in A\}.$$
781:
782: \smallskip\
783:
784: The relation between commutator and strong commutator is
785: clarified by the following lemma. In addition, the third item of
786: the following lemma shows that commutator is a relativization of
787: the notion of commutativity.
788:
789: \smallskip\
790:
791: \textbf{Lemma 2.4}. ([T81], Proposition 4 and its corollary) Let
792: $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular
793: lattice and let $A\subseteq L$. Then
794:
795: (1) $\Gamma(A)\leq \gamma(A)$.
796:
797: (2) If $A$ is finite, then $\Gamma(A)= \gamma(A)$.
798:
799: (3) $\gamma(A)=1$ if and only if all the members of $A$ are
800: mutually commutable.$\heartsuit$
801:
802: \smallskip\
803:
804: We now can present a generalization of Lemma 2.3 by using the tool
805: of commutator. It is easy to see from Lemmas 2.4(2) and (3) that
806: the following lemma degenerates to Lemma 2.3 when $aCb_i$ for all
807: $i\in I$.
808:
809: \smallskip\
810:
811: \textbf{Lemma 2.5}. ([T81], Propositions 5 and 6) Let $\ell
812: =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular lattice and
813: let $A\subseteq L$. Then for any $a\in A$ and $b_i\in A$ $(i\in
814: I)$,
815: $$\Gamma(A)\wedge (a\wedge \vee_{i\in I}b_i)\leq \vee_{i\in I}(a\wedge b_i),$$
816: $$\Gamma(A)\wedge \wedge_{i\in I}(a\vee b_i)\leq a\vee \wedge_{i\in I}b_i.\heartsuit$$
817:
818: \smallskip\
819:
820: Suppose that we want to use the above lemma on a formula of the
821: form $a\wedge \vee_{i\in I}b_i$ or $a\vee \wedge_{i\in I}b_i$ in
822: order to get a local distributivity. In many situations, the
823: elements $a$ and $b_i$ $(i\in I)$ may be very complicated, and the
824: operations $\bot$, $\wedge$ and $\vee$ are involved in them. Then
825: the above lemma cannot be applied directly, and it needs the help
826: of the following
827:
828: \smallskip\
829:
830: \textbf{Lemma 2.6}. Let $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$
831: be an orthomodular lattice and let $A\subseteq L$. Then for any
832: $B\subseteq [A]$ we have $$\Gamma (A)\leq \Gamma (B),$$ where
833: $[A]$ stands for the subalgebra of $\ell$ generated by $A$.
834:
835: \smallskip\
836:
837: \textbf{Proof}. For any $X\subseteq L$, we write
838: $$K(X)=\{b\in L:aCb\ {\rm and}\ (a_1\wedge b)C(a_2\wedge b)\
839: {\rm for\ all}\ a, a_1, a_2 \in X \}$$ Furthermore, we set $A_0=A$
840: and
841: $$A_{i+1} = A_i \cup \{a^{\bot}: a\in A_i\}\cup\{a_1\wedge a_2:
842: a_1, a_2\in A_i\}\ (i=0,1,2,...)$$
843:
844: First, we prove that $K(A_i)=K(A)$ for all $i\geq 0$ by induction
845: on $i$. It is obvious that $K(A_{i+1})\subseteq K(A)$. Conversely,
846: suppose that $b\in K(A)$ and we want to show that $b\in
847: K(A_{i+1})$. It is easy to see that $aCb$ for any $a\in A_{i+1}$.
848: Thus, we only need to demonstrate the following
849:
850: \textit{Claim:} $(a_1\wedge b)C(a_2\wedge b)$ for any $a_1, a_2\in
851: A_{i+1}$.
852:
853: The essential part of the proof of the above claim is the
854: following two cases, and the other cases are clear, or can be
855: treated as iterations of them:
856:
857: Case 1. $a_1\in A_i$, $a_2=c_1\wedge c_2$ and $c_1, c_2\in A_i$.
858: From the induction hypothesis we have $$(a_1\wedge b)C(c_1\wedge
859: b)\ {\rm and}\ (a_1\wedge b)C(c_2\wedge b).$$ This yields
860: $$(a_1\wedge b)C(c_1\wedge b)\wedge (c_2\wedge b)=(c_1\wedge
861: c_2)\wedge b=a_2\wedge b.$$
862:
863: Case 2. $a_1\in A_i$, $a_2=c^{\bot}$ and $c\in A_i$. Then from the
864: induction hypothesis we obtain $(a_1\wedge b)C(c\wedge b)$, and
865: further $(a_1\wedge b)C(c\wedge b)^{\bot}$ by using Lemma 2.1(4).
866: In addition, $(a_1\wedge b)Cb$. This together with Lemma 2.2
867: yields $(a_1\wedge b)Cb\wedge(c\wedge b)^{\bot}$. Note that $cCb$
868: and so $b^{\bot}Cc^{\bot}$. Then by Lemma 2.3 we assert that
869: $$b^{\bot}\vee(c\wedge b)=b^{\bot}\vee c\ {\rm and}\ b\wedge (c\wedge
870: b)^{\bot}=b\wedge c^{\bot}$$. Hence, it follows that $(a_1\wedge
871: b)Cb\wedge c^{\bot}=a_2\wedge b.$
872:
873: We now write $$A_\infty = \cup_{i=0}^{\infty}A_i.$$ Then
874: $$K(A_\infty)=\cap_{i=0}^{\infty}K(A_i)=K(A).$$ It is easy to see
875: that $A\subseteq A_\infty$ is a subalgebra of $\ell$. So,
876: $[A]\subseteq A_\infty$, $$K(A)=K(A_\infty)\subseteq
877: K([A])\subseteq K(B),$$ and $$\Gamma (A)=\vee K(A)\leq \vee
878: K(B)=\Gamma (B).\heartsuit$$
879:
880: \smallskip\
881:
882: As stated in the introduction, the aim of this paper is to develop
883: a theory of computation based on quantum logic. The logical
884: language for a theory of computation has to contain the universal
885: and existential quantifiers, and the two quantifiers are usually
886: interpreted as (infinite) meet and join, respectively. Hence, we
887: should assume that the lattice of the truth values of our quantum
888: logic is complete. A complete orthomodular lattice is an
889: orthomodular lattice $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ in
890: which for any $M\subseteq L,$ both the greatest lower bound
891: $\wedge M$ and the least upper bound $\vee M$ exist.
892:
893: The function of a logic is provide us with a certain reasoning
894: ability, and the implication connective is an intrinsic
895: representative of inference within the logic. Thus each logic
896: should reasonably contain a connective of implication. To make a
897: complete orthomodular lattice available as the set of truth values
898: of quantum logic, we need to define a binary operation, called
899: implication operator, on it such that this operation may serve as
900: the interpretation of implication in this logic. Unfortunately, it
901: is a very vexed problem to define a reasonable implication
902: operator for quantum logic. All implication operators that one can
903: reasonably introduce in an orthomodular lattice are more or less
904: anomalous in the sense that they do not share most of the
905: fundamental properties of the implication in classical logic. This
906: is different from the cases of most weak logics. (For a thorough
907: discussion on the implication problem in quantum logic, see
908: [DC86], Section 3.)
909:
910: A minimal condition for an implication operator $\rightarrow$ is
911: the requirement proposed by G. Birkhoff and J. von Neumann [BN36]:
912: $$a\rightarrow b=1\ {\rm if\ and\ only\ if}\ a\leq b$$ for any $a,b\in L$.
913: Usually in a logic, there are two ways in which implication is
914: introduced. The first one is to treat implication as a derived
915: connective; that is, implication is explicitly defined in terms of
916: other connectives such as negation, conjunction and disjunction.
917: All implications of this kind were found by G. Kalmbach [Ka74],
918: and they are presented by the following:
919:
920: \smallskip\
921:
922: \textbf{Lemma 2.7}. ([Ka74]; see also [Ka83], Theorem 15.3) The
923: orthomodular lattice freely generated by two elements is
924: isomorphic to $2^{4}\times MO2$, where $2$ stands for the Boolean
925: algebra of two elements. The elements of $2^{4}\times MO2$
926: satisfying the Birkhoff-von Neumann requirement are exactly the
927: following five polynomials of two variables:
928: $$a\rightarrow_1 b=(a^{\bot}\wedge b)\vee (a^{\bot}\wedge b^{\bot})\vee (a\wedge (a^{\bot}\vee b)),$$
929: $$a\rightarrow_2 b=(a^{\bot}\wedge b)\vee (a\wedge b)\vee ((a^{\bot}\vee b)\wedge b^{\bot}),$$
930: $$a\rightarrow_3 b=a^{\bot}\vee (a\wedge b),$$
931: $$a\rightarrow_4 b=b\vee (a^{\bot}\wedge b^{\bot}),$$
932: $$a\rightarrow_5 b=(a^{\bot}\wedge b)\vee (a\wedge b)\vee (a^{\bot}\wedge b^{\bot}).\heartsuit$$
933:
934: \smallskip\
935:
936: Obviously, this lemma implies that the above five polynomials are
937: all implication operators definable in orthomodular lattices. It
938: was shown by G. Kalmbach [Ka74, Ka83] that the orthomodular
939: lattice-valued (propositional) logic can be (finitely)
940: axiomatizable by using the modus ponens with
941: implication$\rightarrow_1$ as the only one rule of inference, but
942: the same conclusion does not hold for the other implications
943: $\rightarrow_i$ $(2\leq i\leq 5).$
944:
945: We may also define the material conditional $\rightarrow_0$ in an
946: orthomodular $\ell=<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ by
947: $$a\rightarrow _0 b=a^{\bot}\vee b$$
948: for all $a,b\in L.$ It is easy to see that $\rightarrow_0$ does
949: not fulfil the Birkhoff-von Neumann requirement. On the other
950: hand, the following lemma shows that the five implication
951: operators given in Lemma 2.7 degenerate to the material
952: conditional whenever the two operands are compatible.
953:
954: \smallskip\
955:
956: \textbf{Lemma 2.8}. ([DC86], Theorem 3.2) Let $\ell =<L,\leq
957: ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular lattice. Then for
958: any $a,b\in L$, $$a\rightarrow_i b=a\rightarrow_0 b$$ if and only
959: if $aCb$, where $1\leq i\leq 5.\heartsuit$
960:
961: \smallskip\
962:
963: The second way of defining an implication is to take its truth
964: function as the adjunctor (i.e., residuation) of the truth
965: function of conjunction. Note that in this case the implication is
966: usually not definable from negation, conjunction and disjunction,
967: and it has been treated as a primitive connective. Indeed, L.
968: Herman, E. Marsden and R. Piziak [HMP75] introduced an implication
969: in the style of residuation. Furthermore, the following lemma
970: shows that the five polynomial implication operators
971: $\rightarrow_i$ $(1\leq i\leq 5)$ cannot be defined as the
972: residuation of the conjunction unless $\ell$ is a Boolean algebra.
973:
974: \smallskip\
975:
976: \textbf{Lemma 2.9}. ([DC86], the revised version, page 25) Let
977: $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an orthomodular
978: lattice, and let $1\leq i\leq 5$. Then the following two
979: statements are equivalent:
980:
981: (i) $\ell$ is a Boolean algebra.
982:
983: (ii) the import-export law: for all $a,b\in L$,
984: $$a\wedge b\leq c\ {\rm if\ and\ only\ if}\ a\leq b\rightarrow_i c.\heartsuit$$
985:
986: \smallskip\
987:
988: Among the five orthomodular polynomial implications,
989: $\rightarrow_3$, named the Sasaki-hook, has often been preferred
990: since it enjoys some properties resembling those in intuitionistic
991: logic. The Sasaki-hook was originally introduced by P. D. Finch
992: [Fi70]. For a detailed discussion of the Sasaki-hook, see L.
993: Rom\'{a}n and B. Rumbos [RR91] and L. Rom\'{a}n and R. E. Zuazua
994: [RZ99]. Here we first point out that the Sasaki-hook possesses a
995: modification of residual characterization although it is defined
996: as a polynomial in orthomodular lattice. A weakening of the
997: import-export law is the resulting condition, called compatible
998: import-export law, by restricting the import-export law for any
999: $a,b\in L$ with $aCb$; that is, if $aCb$, then $a\wedge b\leq c$
1000: if and only if $a\leq b\rightarrow c$.
1001:
1002: \smallskip\
1003:
1004: \textbf{Lemma 2.10}. ([T81], Proposition 1 and its corollary;
1005: [DC86], the revised version, page 25) Let $\ell =<L,\leq ,\wedge
1006: ,\vee ,\bot ,0,1>$ be an orthomodular lattice, and let $a,$ $b$,
1007: $c\in L$. Then
1008: $$a\rightarrow b=\vee \{x:xCa\ {\rm and}\ x\wedge a\leq b\}.$$
1009: Moreover, among the five implications $\rightarrow_i$ $(1\leq
1010: i\leq 5)$, the Sasaki-hook $\rightarrow_3$ is the only one
1011: satisfying the compatible import-export law.$\heartsuit$
1012:
1013: \smallskip\
1014:
1015: Our mathematical reasoning frequently require that implication
1016: relation is preserved by conjunction and disjunction. Also, the
1017: negation is needed to be compatible with implication in the sense
1018: that the negation can reverse the direction of implication. And,
1019: to warrant the validity of a chain of inferences, the transitivity
1020: of implication is required. However, this is not the case in
1021: general if we are working in an orthomodular lattice. Fortunately,
1022: if we adopt the Sasaki-hook, then these properties of implication
1023: can be recovered by attaching a certain commutator.
1024:
1025: \smallskip\
1026:
1027: \textbf{Lemma 2.11.} Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
1028: ,0,1>$ be an orthomodular lattice. Then
1029:
1030: (1) for any $a_{i},b_{i}\in L$ $(i=1,...,n),$ let
1031: $X=\{a_{1},...,a_{n}\}\cup \{b_{1},...,b_{n}\},$
1032: $$\Gamma
1033: (X)\wedge \wedge _{i=1}^{n}(a_{i}\rightarrow _{3}b_{i})\leq \wedge
1034: _{i=1}^{n}a_{i}\rightarrow _{3}\wedge _{i=1}^{n}b_{i},$$
1035: $$\Gamma (X)\wedge \wedge _{i=1}^{n}(a_{i}\rightarrow
1036: _{3}b_{i})\leq \vee _{i=1}^{n}a_{i}\rightarrow _{3}\vee
1037: _{i=1}^{n}b_{i}.$$
1038:
1039: (2) for any $a,b\in L$, $$\Gamma (a,b)\wedge (a\rightarrow
1040: _{3}b)\leq b^{\perp }\rightarrow _{3}a^{\perp }.$$
1041:
1042: (3) for any $a,b,c\in L,$
1043: $$\Gamma (a,b,c)\wedge (a\rightarrow _{3}b)\wedge (b\rightarrow
1044: _{3}c)\leq a\rightarrow _{3}c.$$
1045:
1046: \smallskip\
1047:
1048: \textbf{Proof.} (1) We only prove the first inequality, and the
1049: proof of the second is similar. With Lemmas 2.5 and 2.6 we obtain
1050: $$\wedge
1051: _{i=1}^{n}a_{i}\rightarrow _{3}\wedge _{i=1}^{n}b_{i}=(\wedge
1052: _{i=1}^{n}a_{i})^{\perp }\vee (\wedge _{i=1}^{n}a_{i}\wedge \wedge
1053: _{i=1}^{n}b_{i})$$
1054: $$=\vee _{i=1}^{n}a_{i}{}^{\perp }\vee \wedge
1055: _{i=1}^{n}(a_{i}\wedge b_{i})$$
1056: $$\geq \Gamma (X)\wedge \wedge
1057: _{i=1}^{n}(\vee _{j=1}^{n}a_{j}{}^{\perp }\vee (a_{i}\wedge
1058: b_{i}))$$
1059: $$\geq \Gamma (X)\wedge \wedge _{i=1}^{n}(a_{i}{}^{\perp }\vee
1060: (a_{i}\wedge b_{i}))$$
1061: $$=\Gamma (X)\wedge \wedge
1062: _{i=1}^{n}(a_{i}\rightarrow b_{i}).$$
1063:
1064: (2) First, we note that $a\wedge b,$ $a^{\perp }\wedge b,$ $
1065: a^{\perp }\wedge b^{\perp }\leq b\vee (a^{\perp }\wedge b^{\perp
1066: })=b^{\perp }\rightarrow _{3}a^{\perp }.$ Thus,
1067: $$\Gamma (a,b)=(a\wedge b)\vee (a\wedge b^{\perp })\vee (
1068: a^{\perp }\wedge b)\vee (a^{\perp }\wedge b^{\perp })$$
1069: $$\leq (b^{\perp }\rightarrow _{3}a^{\perp })\vee (a\wedge b^{\perp
1070: }),$$ and furthermore with Lemmas 2.5 and 2.6 we have
1071: $$\Gamma (a,b)\wedge (a\rightarrow _{3}b)=\Gamma (a,b)\wedge
1072: (a^{\perp }\vee (a\wedge b))$$
1073: $$\leq \Gamma (a,b)\wedge (a^{\perp }\vee b)$$
1074: $$=\Gamma (a,b)\wedge \Gamma (a,b)\wedge (a^{\perp }\vee b)$$
1075: $$\leq \Gamma (a,b)\wedge \lbrack (b^{\perp }\rightarrow
1076: _{3}a^{\perp })\vee (a\wedge b^{\perp })]\wedge (a^{\perp }\vee
1077: b)$$
1078: $$\leq \ [(b^{\perp }\rightarrow _{3}a^{\perp })\wedge (a^{\perp
1079: }\vee b)]\vee \lbrack (a\wedge b^{\perp })\wedge (a^{\perp }\vee
1080: b)]$$
1081: $$\leq (b^{\perp }\rightarrow _{3}a^{\perp })\vee \lbrack (a\wedge
1082: b^{\perp })\wedge (a^{\perp }\vee b)].$$ Note that $(a\wedge
1083: b^{\perp })^{\perp }=a^{\perp }\vee b$ and $(a\wedge b^{\perp
1084: })\wedge (a^{\perp }\vee b)=0.$ Then
1085: $$\Gamma (a,b)\wedge (a\rightarrow _{3}b)\leq b^{\perp }\rightarrow
1086: _{3}a^{\perp }.$$
1087:
1088: (3) Again, we use Lemmas 2.5 and 2.6. This enables us to assert
1089: that
1090: $$\Gamma
1091: (a,b,c)\wedge (a\rightarrow _{3}b)\wedge (b\rightarrow
1092: _{3}c)=\Gamma (a,b,c)\wedge (a^{\perp }\vee (a\wedge b))\wedge
1093: (b^{\perp }\vee (b\wedge c)) $$
1094: $$\leq \Gamma (a,b,c)\wedge
1095: ([a^{\perp }\wedge (b^{\perp }\vee (b\wedge c))]\vee \lbrack
1096: (a\wedge b)\wedge (b^{\perp }\vee (b\wedge c))])$$
1097: $$\leq \Gamma
1098: (a,b,c)\wedge (a^{\perp }\vee \lbrack (a\wedge b)\wedge (b^{\perp
1099: }\vee (b\wedge c))]).$$
1100:
1101: We note that $\Gamma (a,b,c)Ca^{\perp }$ and $$\Gamma
1102: (a,b,c)C[(a\wedge b)\wedge (b^{\perp }\vee (b\wedge c))]).$$ Then
1103: $$\Gamma (a,b,c)\wedge (a\rightarrow _{3}b)\wedge (b\rightarrow
1104: _{3}c)\leq (\Gamma (a,b,c)\wedge a^{\perp })\vee (\Gamma
1105: (a,b,c)\wedge \lbrack (a\wedge b)\wedge (b^{\perp }\vee (b\wedge
1106: c))])$$
1107: $$\leq a^{\perp }\vee (\Gamma (a,b,c)\wedge \lbrack (a\wedge
1108: b)\wedge (b^{\perp }\vee (b\wedge c))])$$
1109: $$\leq a^{\perp }\vee
1110: \lbrack (a\wedge b)\wedge b^{\perp }]\vee \lbrack (a\wedge
1111: b)\wedge (b\wedge c)]$$
1112: $$=a^{\perp }\vee \lbrack (a\wedge b)\wedge
1113: (b\wedge c)]$$
1114: $$\leq a^{\perp }\vee (a\wedge c)$$
1115: $$=a\rightarrow _{3}c.\heartsuit$$
1116:
1117: \smallskip\
1118:
1119: For simplicity of presentation, we finally introduce an
1120: abbreviation. For each implication operator $\rightarrow$, the
1121: bi-implication operator on $\ell $ is defined as follows:
1122: $$a\leftrightarrow b\stackrel{def}{=}(a\rightarrow b)\wedge
1123: (b\rightarrow a)$$ for any $a,b\in L.$
1124:
1125: \smallskip\
1126:
1127: \textbf{2.2. The Language of Quantum Logic}
1128:
1129: \smallskip\
1130:
1131: In this subsection we present the syntax of quantum logic. Given a
1132: complete orthomodular lattice $\ell =<L,\leq ,\wedge ,\vee ,\bot
1133: ,0,1>.$ We require that the language of an $\ell -$valued
1134: (quantum) logic possesses a nullary connective $ \mathbf{a}$ for
1135: each $a\in L$ as well as three other primitive connectives: an
1136: unary one $\lnot $ (negation) and two binary ones $\wedge $
1137: (conjunction), $\rightarrow$ (implication). The language also has
1138: a primitive quantifier $\forall $ (universal quantifier).
1139:
1140: It deserves an explanation for our design decision of choosing
1141: implication as a primitive connective. In the sequel, many results
1142: only need to suppose that the implication operator satisfies the
1143: Birkhoff-von Neumann requirement. It is known that there are five
1144: polynomials fulfilling the Birkhoff-von Neumann requirement. If we
1145: treated implication as a derived connective defined in terms of
1146: negation, conjunction and disjunction, then it would be necessary
1147: to assume five different connectives of implication in our logical
1148: language. This would often complicate our presentation very much.
1149: On the other hand, in some cases, the Birkhoff-von Neumann
1150: condition is not enough and it requires the implication operator
1151: to be the Sasaki-hook. So, we decide to use implication as a
1152: primitive connective, and specify it when needed.
1153:
1154: The syntax of $\ell-$valued logic is defined in a familiar way; we
1155: omit its details. To simplify the notations in what follows, it is
1156: necessary to introduce several derived formulas:
1157:
1158: (i) $\varphi \vee \psi \stackrel{def}{=}\lnot (\lnot \varphi
1159: \wedge \lnot \psi );$
1160:
1161: (ii) $\varphi \leftrightarrow \psi \stackrel{def}{=}(\varphi
1162: \rightarrow \psi )\wedge (\psi \rightarrow \varphi );$
1163:
1164: (iii) $(\exists x)\varphi \stackrel{def}{=}\lnot (\forall x)\lnot
1165: \varphi ;$
1166:
1167: (iv) $A\subseteq B\stackrel{def}{=}(\forall x)(x\in A\rightarrow
1168: x\in B);$ and
1169:
1170: (v) $A\equiv B\stackrel{def}{=}(A\subseteq B)\wedge (B\subseteq
1171: A).$
1172:
1173: Suppose that $\Delta$ is a finite set of formulas. The commutator
1174: of $\Delta$ is defined to be
1175: $$\gamma(\Delta)\stackrel{def}{=}\vee\{\wedge_{\varphi\in\Delta}
1176: \varphi^{f(\varphi)}:f\in \{1,-1\}^{\Delta}\},$$ where
1177: $\varphi^{1},\ \varphi^{-1}$ express $\varphi$ and $\lnot\varphi,$
1178: respectively. It is obvious that the above formula is the
1179: counterpart of Definition 2.1(1) in the language of our quantum
1180: logic.
1181:
1182: \smallskip\
1183:
1184: \textbf{2.3. The Algebraic Semantics of Quantum Logic}
1185:
1186: \smallskip\
1187:
1188: We now turn to give the semantics of quantum logic. There are
1189: several different versions of semantics for quantum logic; for
1190: example, quantum logic enjoys a semantics in the Kripke style
1191: [DC86; RS00]. What concerns us here is its algebraic semantics.
1192: Assume that $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be an
1193: orthomodular lattice equipped with additionally a binary operation
1194: $\rightarrow$ over it. The operation $\rightarrow$ is required to
1195: be suited to serve as the truth function of implication
1196: connective. According to our explanation of the connective of
1197: implication in the last subsection, we leave the operation
1198: $\rightarrow$ unspecified but suppose that it satisfies the
1199: Birkhoff-von Neumann requirement. An $\ell -$valued interpretation
1200: is an interpretation in which every predicate symbol is associated
1201: with an $ \ell -$valued relation, i.e., a mapping from the product
1202: of some copies of the discourse universe into $L,$ where the
1203: number of copies is exactly the arity of the predicate symbol. The
1204: other items in $\ell-$valued logical language are interpreted as
1205: usual. For every (well-formed) formula $\varphi ,$ its truth value
1206: $\lceil \varphi \rceil$ is assumed in $L,$ and the truth valuation
1207: rules for logical and set-theoretical formulas are given as
1208: follows:
1209:
1210: (i) $\lceil \mathbf{a}\rceil =a;$
1211:
1212: (ii) $\lceil \lnot \varphi \rceil =\lceil \varphi \rceil ^{\bot
1213: };$
1214:
1215: (iii) $\lceil \varphi \wedge \psi \rceil =\lceil \varphi \rceil
1216: \wedge \lceil \psi \rceil ;$
1217:
1218: (iv) $\lceil \varphi \rightarrow \psi \rceil =\lceil \varphi
1219: \rceil \rightarrow \lceil \psi \rceil ;$
1220:
1221: (v) if $U$ is the universe of discourse, then
1222: $$\lceil (\forall x)\varphi (x)\rceil =\wedge _{u\in U}\lceil
1223: \varphi (u)\rceil ;$$ and
1224:
1225: (vi) $\lceil x\in A\rceil =A(x),$ where $A$ is a set constant
1226: (unary predicate symbol) and it is interpreted as a mapping, also
1227: denoted as $A,$ from the universe into $L,$ i.e., an $ \ell
1228: -$valued set (more exactly, an $\ell -$valued subset of the
1229: universe).
1230:
1231: Note that in the above truth valuation rules $\wedge $ and $\vee $
1232: in the left-hand side are connectives in quantum logic whereas
1233: $\wedge $ and $\vee $ in the right-hand side stand for operations
1234: in the orthomodular lattice $ \ell $ of truth values. Also, the
1235: symbol $\rightarrow$ in the left-hand side of (iv) is a connective
1236: in the language of quantum logic, but the symbol $\rightarrow$ in
1237: the right-hand side of (iv) is the binary operation attached to
1238: $\ell$ that is explained at the beginning of this subsection.
1239:
1240: As we claimed in the introduction, quantum logic will act as our
1241: meta-logic in the theory of computation developed in this paper.
1242: Then we still have to introduce several meta-logical notions for
1243: quantum logic. For every orthomodular lattice $\ell =<L,\leq
1244: ,\wedge ,\vee ,\bot ,0,1>,$ if $\Gamma $ is a set of formulas and
1245: $\varphi $ a formula, then $\varphi $ is a semantic consequence of
1246: $\Gamma $ in $\ell -$valued logic, written $\Gamma \stackrel{\ell
1247: }{\models }\varphi ,$ whenever $$\wedge _{\psi \in \Gamma }\lceil
1248: \psi \rceil \leq \lceil \varphi \rceil $$ for all $\ell -$valued
1249: interpretations. In particular, $\stackrel{\ell }{\models }\varphi
1250: $ means that $\phi \stackrel{\ell }{\models }\varphi ,$ i.e.,
1251: $\lceil \varphi \rceil =1$ always holds for every $\ell -$valued
1252: interpretation; in other words, $1$ is the unique designated truth
1253: value in $\ell .$ Furthermore, if $\Gamma \stackrel{ \ell
1254: }{\models }\varphi $ (resp. $\stackrel{\ell }{\models }\varphi $)
1255: for all orthomodular lattice $\ell $ then we say that $\varphi $
1256: is a semantic consequence of $\Gamma $ (resp. $\varphi $ is valid)
1257: in quantum logic and write $\Gamma \models \varphi $ (resp.
1258: $\models \varphi $).
1259:
1260: We here are not going to give a detailed exposition on quantum
1261: logic, but would like to point out that quantum logic gives rise
1262: to many counterexamples to some meta-logical properties which hold
1263: for classical logic and for a large class of weaker logics; for
1264: example, M. L. Dalla Chiara [DC81] showed that a minimal version
1265: of quantum logic fails to enjoy the Lindenbaum property, and J.
1266: Malinowski [Ma90] found that the deduction theorem fails in
1267: quantum logic and some of its variants.
1268:
1269: \smallskip\
1270:
1271: \textbf{2.4. The Operations of Quantum Sets}
1272:
1273: \smallskip\
1274:
1275: Beside the language of quantum logic introduced in Section 2.2, we
1276: will also need some notations such as $\in $ (membership) from
1277: set-theoretical language in our study of computing theory based on
1278: quantum logic. As mentioned in the introduction, a theory of
1279: quantum sets has already been developed by G. Takeuti [T81]. A
1280: careful review of quantum set theory is out of the scope of the
1281: present paper. What mainly concerned G. Takeuti [T81] is how some
1282: axioms of classical set theory could be modified so that they will
1283: holds in the framework of quantum logic. In other words, he tried
1284: to clarify the relation of quantum set theory with the classical
1285: mathematics. Here, we instead propose some operations of
1286: $\ell-$valued sets and also introduce several notations for
1287: $\ell-$valued sets. These are needed in the subsequent sections.
1288: We write $L^{X}$ for the set of all $\ell-$valued subsets of $X$,
1289: i.e., all mappings from $X$ into $L$. For any non-empty set $X,$
1290: if $x\in X$ and $\lambda \in L-\{0\},$ then $x_\lambda $ is
1291: defined to be a mapping from $X$ into $L$ such that
1292: $$x_\lambda (x^{\prime })=\left\{
1293: \begin{array}{c}
1294: \lambda \ {\rm if }\ x^{\prime }=x, \\
1295: 0\ {\rm otherwise,}
1296: \end{array}
1297: \right.$$ and it is often called an $\ell-$valued point in $X.$ We
1298: write $p_\ell(X)$ for the set of all $\ell-$valued points in $X;$
1299: that is, $$p_\ell(X)=\{x_\lambda :x\in X\ {\rm and}\ \lambda \in
1300: L-\{0\}\}.$$ For each $e=x_\lambda \in p_\ell(X),$ $x$ is called
1301: the support of $e$ and denoted $s(e),$ and $\lambda $ is called
1302: the height of $e$ and written $h(e).$ In particular, an
1303: $\ell-$valued point of height $1$ is always identified with its
1304: support. The predicate $\in$ can be extended to a predicate
1305: between $\ell-$valued points and $\ell-$valued sets in a natural
1306: way: $$x_{\lambda}\in A\stackrel{def}{=}x_{\lambda}\subseteq A.$$
1307: Then it is easy to see that $$\lceil x_{\lambda}\in A\rceil =
1308: \lambda\rightarrow A(x)$$ for any $x\in X$, $\lambda\in L$ and
1309: $A\in L^{X}$, where $\rightarrow$ is the implication operator
1310: under consideration. For any $A\subseteq X$, its characteristic
1311: function is a mapping from $X$ into the Boolean algebra
1312: $\mathbf{2}=\{0,1\}$ of two elements, and so it can also be seen
1313: as a mapping from $X$ into $L$, namely, an $\ell-$valued subset of
1314: $X$. We will identify $A$ with its characteristic function. For
1315: any $a\in L$ and $A,B\in L^{X}$, we define all of the scalar
1316: product $aA,$ complement $A^{c}$, intersection $A\cap B$ and union
1317: $A\cup B$ to be $\ell-$valued subsets of $X$ and for all $x\in X$,
1318:
1319: \smallskip\
1320:
1321: (i) $x\in aA\stackrel{def}{=}\mathbf{a}\wedge (x\in A);$
1322:
1323: (ii) $x\in A^{c}\stackrel{def}{=}\lnot (x \in A);$
1324:
1325: (iii) $x\in A\cap B\stackrel{def}{=}(x\in A)\wedge (x\in B);$
1326:
1327: (iv) $x\in A\cup B\stackrel{def}{=}(x\in A)\vee (x\in B).$
1328:
1329: \smallskip\
1330:
1331: From the truth valuation rules and the definition of derived
1332: formulas in the $\ell-$valued logical and set-theoretical
1333: language, we know that for all $x\in X$,
1334:
1335: \smallskip\
1336:
1337: (i') $(aA)(x)=a \wedge A(x);$
1338:
1339: (ii') $(A^{c})(x)=A(x)^{\bot};$
1340:
1341: (iii') $(A\cap B)(s)=A(s)\wedge B(s);$ and
1342:
1343: (iv') $(A\cup B)(s)=A(s)\vee B(s).$
1344:
1345: \smallskip\
1346:
1347: It is easy to see that in the domain of $\ell-$valued sets the
1348: intersection and union operations are idempotent, commutative and
1349: associative, and they have $X$ and $\phi$, respectively as their
1350: unit elements. The intersection and union together with the
1351: complement satisfy the De Morgan law, but the distributivity of
1352: intersection over union or union over intersection is no longer
1353: valid. Clearly, the laws for operations of $\ell-$valued sets are
1354: essentially determined by the algebraic properties of the lattice
1355: $\ell$ of truth values.
1356:
1357: Assume that $X$ and $Y$ are two non-empty sets, and
1358: $h:X\longrightarrow Y$ is a mapping. For any $A\in L^{X}$, its
1359: image $h(A)$ under $h$ is defined by
1360: $$y\in h(A)\stackrel{def}{=}(\exists x\in X)(y=f(x)\wedge x\in
1361: A,$$ and for any $B\in L^{Y}$, its pre-image $h^{-1}(B)$ under $h$
1362: is defined by $$x\in h^{-1}(B)\stackrel{def}{=}h(x)\in B.$$ The
1363: defining equations of $h(A)$ and $h^{-1}(B)$ may be rewritten,
1364: respectively, as follows: $$h(A)(y)=\vee \{A(X):x\in X\ {\rm and}\
1365: f(x)=y\},\ {\rm and}$$ $$h^{-1}(B)(x)=B(h(x)).$$
1366:
1367: \smallskip\
1368:
1369: \textbf{Lemma 2.12.} Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
1370: ,0,1>$ be an orthomodular lattice, let $\rightarrow $ enjoy the
1371: Birkhoff-von Neumann requirement, and let $h:X\rightarrow Y$ be a
1372: mapping. Then for any $A,B\in L^{Y},$
1373: $$\stackrel{\ell }{\models} A\equiv B\rightarrow h^{-1}(A)\equiv h^{-1}(B).$$
1374:
1375: \smallskip\
1376:
1377: \textbf{Proof.} $$\lceil h^{-1}(A)\equiv h^{-1}(B)\rceil =\wedge
1378: _{x\in X}(h^{-1}(A)(x)\longleftrightarrow h^{-1}(B)(x))$$
1379: $$=\wedge _{x\in X}(A(h(x))\longleftrightarrow B(h(x)))$$
1380: $$\geq \wedge _{y\in Y}(A(y)\longleftrightarrow B(y))$$
1381: $$=\lceil A\equiv B\rceil .\heartsuit$$
1382:
1383: \smallskip\
1384:
1385: To conclude this section, we introduce the notion of $\ell-$valued
1386: language as well as some operations of $\ell-$valued languages.
1387: Suppose that $\Sigma$ is an alphabet; that is, a finite nonempty
1388: set (of input symbols). We write $\Sigma^{*}$ for the set of
1389: strings over $\Sigma$:
1390: $$\Sigma^{*}=\cup_{n=0}^{\infty} \Sigma^{n}.$$
1391: An $\ell-$valued language over $\Sigma$ is defined to be an
1392: $\ell-$valued subset of $\Sigma^{\ast}$. Thus, the set of
1393: $\ell-$valued languages over $\Sigma$ is exactly
1394: $L^{\Sigma^{\ast}}$. Let $A,B\in L^{\Sigma ^{\ast }}$ be two
1395: $\ell-$valued subsets of $\Sigma^{*}$. Then we define the
1396: concatenation $A\cdot B$ of $A$ and $B$ and the Kleene closure
1397: $A^{\ast }\in L^{\Sigma ^{\ast }}$ of $A$ as follows: for any
1398: $s\in \Sigma ^{\ast },$
1399:
1400: \smallskip\
1401:
1402: (v) $s\in A\cdot B\stackrel{def}{=}(\exists u,v\in
1403: \Sigma^{\ast})(s=uv\wedge u\in A\wedge v\in B);$
1404:
1405: (vi) $s\in A^{\ast }\stackrel{def}{=}(\exists n\geq 0)(\exists
1406: s_{1},...,s_{n}\in \Sigma ^{\ast }(s=s_{1}...s_{n}\wedge \wedge
1407: _{i=1}^{n}(s_{i}\in A)).$
1408:
1409: \smallskip\
1410:
1411: The above defining equations can also be translated to the
1412: following two formulas in the lattice of truth values by employing
1413: the truth valuation rules: for every $s\in \Sigma^{*}$,
1414:
1415: \smallskip\
1416:
1417: (v') $(A\cdot B)(s)=\vee \{A(u)\wedge B(v):u,v\in \Sigma ^{\ast }\
1418: {\rm and}\ s=uv\};$
1419:
1420: (vi') $A^{\ast }(s)=\vee \{\wedge _{i=1}^{n}A(s_{i}):n\geq
1421: 0,s_{1},...,s_{n}\in \Sigma ^{\ast }\ {\rm and}\
1422: s=s_{1}...s_{n}\}$.
1423:
1424: \smallskip\
1425:
1426: It is easy to demonstrate that if the meet $\wedge$ is
1427: distributive over the join $\vee$ in $\ell$ (in other words,
1428: $\ell$ is a Boolean algebra), then we have
1429: $$A^{*}=\cup_{n=0}^{\infty}A^{n}$$ where
1430: $$\left\{
1431: \begin{array}{c}
1432: A^{0}=\{\varepsilon \}, \\
1433: A^{n+1}=A^{n}\cdot A\ {\rm for\ all }\ n\geq 0.
1434: \end{array}
1435: \right. $$
1436:
1437: \smallskip\
1438:
1439: \textbf{3. Orthomodular Lattice-Valued Automata}
1440:
1441: \smallskip\
1442:
1443: For convenience we first recall some basic notions in classical
1444: automata theory. Let $\Sigma $ be a finite input alphabet whose
1445: elements are called input symbols or labels. Then a
1446: nondeterministic finite automaton (NFA for short) over $\Sigma $
1447: is a quadruple
1448: $$\Re =<Q,I,T,E>$$
1449: in which:
1450:
1451: (i) $Q$ is a finite set whose elements are called states;
1452:
1453: (ii) $I\subseteq Q$ and states in $I$ are said to be initial;
1454:
1455: (iii) $T\subseteq Q$ and states in $T$ are said to be terminal;
1456: and
1457:
1458: (iv) $E\subseteq Q\times \Sigma \times E,$ and each $(p,\sigma
1459: ,q)\in E$ is called a transition in (or an edge of) $\Re $ and it
1460: means that input $\sigma $ makes state $p$ evolves to $q.$
1461:
1462: An NFA is said to be deterministic if $I$ is a singleton, and for
1463: any $p$ in $Q$ and $\sigma$ in $\Sigma$, there is exactly one $q$
1464: in $Q$ such that $(p,\sigma,q)\in E$. Thus, the transition
1465: relation $E$ in a deterministic finite automaton (DFA, for short)
1466: may be seen as a mapping from $Q\times \Sigma $ into $Q,$ and it
1467: is called the transition function.
1468:
1469: A path in $\Re $ is a finite sequence of the form
1470: $$c=q_0\sigma_1q_1...q_{k-1}\sigma _kq_k$$
1471: such that $(q_i,\sigma _{i+1},q_{i+1})\in E$ for each $i<k.$ In
1472: this case, the sequence $\sigma _1...\sigma _k$ is called the
1473: label of $c.$ A path $c=q_0\sigma _1q_1...q_{k-1}\sigma _kq_k$ is
1474: said to be successful if $q_0\in I$ and $q_k\in T.$ The language
1475: accepted by an automaton $\Re $ is the set of labels of all
1476: successful paths in $\Re .$ Let
1477: $$A\subseteq \Sigma ^{*}=\cup _{n=0}^\infty \Sigma ^n.$$
1478: Then $A$ is said to be regular if there is an automaton $\Re $
1479: over $\Sigma $ such that $A$ is the language accepted by $\Re .$
1480:
1481: The notion of orthomodular lattice-valued automata is a natural
1482: generalization of NFAs. Let $\ell =<L,\leq ,\wedge ,\vee ,\bot
1483: ,0,1>$ be an orthomodular lattice, and let $\Sigma $ be a finite
1484: alphabet. Then an $\ell -$valued (quantum) automaton over $\Sigma
1485: $ is a quadruple $$\Re =<Q,I,T,\delta >$$ where:
1486:
1487: (i) $Q$ is the same as in an NFA;
1488:
1489: (ii) $I$ is an $\ell-$valued subset of $Q$; that is, a mapping
1490: from $Q$ into $L$. For each $q\in Q$, $I(q)$ indicates the truth
1491: value (in the underlying quantum logic) of the proposition that
1492: $q$ is an initial state;
1493:
1494: (iii) $T$ is also an $\ell-$valued subset of $Q$, and for every
1495: $q\in Q$, $T(q)$ expresses the truth value (in our quantum logic)
1496: of the proposition that $q$ is terminal; and
1497:
1498: (iv) $\delta $ is an $\ell -$valued subset of $Q\times \Sigma
1499: \times Q$; that is, a mapping from $Q\times \Sigma \times Q$ into
1500: $L,$ and it is called the $\ell -$valued (quantum) transition
1501: relation of $\Re .$ Intuitively, $\delta$ is an $\ell -$valued
1502: (ternary) predicate over $ Q,\Sigma $ and $Q$, and for any $p,q\in
1503: Q$ and $\sigma\in \Sigma$, $\delta (p,\sigma ,q)$ stands for the
1504: truth value (in quantum logic) of the proposition that input
1505: $\sigma $ causes state $p$ to become $q.$
1506:
1507: The propositions of the form
1508: $$"q\ {\rm is\ an\ initial\ state}",\ {\rm written}\ "q\in I",$$
1509: $$"q\ {\rm is\ a\ terminal\ state}",\ {\rm written}\ "q\in T",$$
1510: and
1511: $$"{\rm input}\ \sigma\ {\rm causes\ state}\ p\ {\rm to\ become}\ q,\
1512: {\rm according\ to\ the\ specification}$$ $${\rm given\ by}\
1513: \delta,"\ {\rm written}\ "p\stackrel{\delta,\sigma
1514: }{\longrightarrow }q"$$ are assumed to be atomic propositions in
1515: our logical language designated for describing $\ell-$valued
1516: automata $\Re$. The truth values of the above three propositions
1517: are respectively $I(q)$, $T(q)$ and $\delta(p,\sigma,q)$. The set
1518: of these atomic propositions is denoted $atom(\Re)$. Formally, we
1519: have
1520: $$atom(\Re)=\{"q\in I":q\in Q\}\cup \{"q\in T":q\in Q\}\cup \{"p\stackrel{\delta,\sigma }
1521: {\longrightarrow }q":p,q\in Q\ {\rm and}\ \sigma\in \Sigma\}.$$
1522:
1523: We write $\mathbf{A}(\Sigma ,\ell )$ for the (proper) class of all
1524: $\ell -$valued automata over $\Sigma.$
1525:
1526: Before defining the concept of recognizability for $\ell -$valued
1527: automata, we need to introduce some auxiliary notions and
1528: notations. We set
1529: $$T(Q,\Sigma )=(Q\Sigma )^{*}Q=\cup _{n=0}^\infty [(Q\Sigma )^n\ Q];$$
1530: that is, the set of all alternative sequences of states and labels
1531: beginning at a state and also ending at a state. For any
1532: $c=q_0\sigma _1q_1...q_{k-1}\sigma _kq_k\in T(Q,\Sigma ),$ the
1533: length of $c$ is defined to be $k$ and denoted by $|c|,$ $q_0$ is
1534: the beginning of $c$ and denoted by $b(c),$ $q_k$ is the end of
1535: $c$ and denoted by $e(c),$ and sequence $s=\sigma _1...\sigma _k$
1536: is called the label of $c$ and denoted by $lb(c).$
1537:
1538: Let $\Re \in \mathbf{A}(\Sigma ,\ell )$ be an $\ell-$valued
1539: automaton over $\Sigma$. Then the $\ell -$valued (unary) predicate
1540: $path_\Re $ on $T(Q,\Sigma )$ is defined as $path_\Re \in
1541: L^{T(Q,\Sigma )}$ (the set of all mappings from $ T(Q,\Sigma )$
1542: into $L$):
1543: $$path_\Re (c)\stackrel{def}{=}\wedge _{i=0}^{k-1}[(q_i,\sigma
1544: _{i+1},q_{i+1})\in \delta ]$$ for every $c=q_0\sigma
1545: _1q_1...q_{k-1}\sigma _kq_k\in T(Q,\Sigma ).$
1546:
1547: In intuition, the truth value of the proposition that $c=q_0\sigma
1548: _1q_1...q_{k-1}\sigma _kq_k$ is a path in $\Re $ is
1549: $$\lceil path_\Re (c)\rceil =\wedge _{i=0}^{k-1}\delta (q_i,\sigma _{i+1},q_{i+1}).$$
1550: Note the difference between the symbols $\wedge$ in the above two
1551: equations: the former is a logical connective, whereas the latter
1552: is an operation on the lattice of truth values.
1553:
1554: Now, we are ready to define one of the key notions in this paper,
1555: namely, recognizability for $\ell -$valued automata. It will be
1556: seen that the defining equation of $\ell-$valued recognizability
1557: is the same as that in the classical theory of automata. The
1558: essential difference between the quantum recognizability and the
1559: corresponding classical notion implicitly resides in their truth
1560: values.
1561:
1562: \smallskip\
1563:
1564: \textbf{Definition 3.1}. Let $\Re \in \mathbf{A}(\Sigma ,\ell )$.
1565: Then the $\ell -$valued (unary) recognizability predicate $rec_\Re
1566: $ on $\Sigma ^{*}$ is defined as $rec_\Re \in L^{\Sigma ^{*}}:$
1567: for every $s\in \Sigma ^{*},$
1568: $$rec_\Re (s)\stackrel{def}{=}(\exists c\in
1569: T(Q,\Sigma ))(b(c)\in I\wedge e(c)\in T\wedge lb(c)=s\wedge
1570: path_\Re (c)).$$ In other words, the truth value of the
1571: proposition that $s$ is recognizable by $ \Re $ is given by
1572: $$\lceil rec_\Re (s)\rceil
1573: =\vee \{I(b(c))\wedge T(e(c))\wedge\lceil path_\Re (c)\rceil :c\in
1574: T(Q,\Sigma )\ {\rm and}\ lb(c)=s\}.$$
1575:
1576: \smallskip\
1577:
1578: We note that $rec_\Re $ is defined above as an $\ell -$valued
1579: unary predicate on $\Sigma ^{*},$ so it may also be seen as an
1580: $\ell -$valued subset of $\Sigma ^{*};$ that is, a mapping
1581: $rec_\Re :\Sigma ^{*}\rightarrow L$ with $rec_\Re (s)=\lceil
1582: rec_\Re (s)\rceil $ for all $s\in \Sigma ^{*}.$
1583:
1584: As a straightforward generalization of regular language, we can
1585: also define regularity for $\ell-$valued languages.
1586:
1587: \smallskip\
1588:
1589: \textbf{Definition 3.2}. The $\ell -$valued (unary) regularity
1590: predicate $Reg_\Sigma $ on $L^{\Sigma ^{*}}$ (the set of all $\ell
1591: -$valued subsets of $\Sigma ^{*}$ ) is defined as $Reg_\Sigma \in
1592: L^{(L^{\Sigma ^{*}})}:$ for each $A\in L^{\Sigma ^{*}},$
1593: $$Reg_\Sigma (A)\stackrel{def}{=}(\exists \Re \in \mathbf{A}(\Sigma
1594: ,\ell ))(A\equiv rec_\Re ).$$ Thus, the truth value of the
1595: proposition that $A$ is regular is
1596: $$\lceil Reg_\Sigma (A)\rceil =\vee \{\lceil A\equiv rec_\Re \rceil
1597: :\Re \in \mathbf{A}(\Sigma ,\ell )\}.$$
1598:
1599: It should be noted that the (automaton) variable $\Re $ bounded by
1600: the existential quantifier in the right-hand side of the defining
1601: formula of $ Reg_\Sigma $ ranges over the proper class
1602: $\mathbf{A}(\Sigma ,\ell ).$ Some readers who are familiar with
1603: axiomatic set theory may worry about that this definition will
1604: cause a certain set-theoretical difficulty, but we stay well away
1605: from anything genuinely problematic. Indeed, for any $\ell
1606: -$valued automaton $\Re =<Q,I,T,\delta >,$ there is a bijection
1607: $\varsigma :Q\rightarrow |Q|$ (the cardinality of $Q$)
1608: $=\{0,1,...,|Q|-1\}$ and we can construct a new $\ell -$valued
1609: automaton $$\varsigma (\Re )=<|Q|,\varsigma (I),\varsigma
1610: (T),\varsigma (\delta )>$$ where $$\varsigma (\delta )(m,\sigma
1611: ,n)=\delta (\varsigma ^{-1}(m),\sigma ,\varsigma ^{-1}(n))$$ for
1612: any $m,n\in |Q|$ and $\sigma \in \Sigma .$ It is easy to see that
1613: $rec_\Re =rec_{\varsigma (\Re )}.$ Then in Definition 3.2 we may
1614: only require that the variable $\Re $ bounded by the existential
1615: quantifier ranges over all $ \ell -$valued automata whose state
1616: sets are subsets of $\omega $ (the set of all non-negative
1617: integers) and the class of all $\ell -$valued automata with
1618: subsets of $\omega $ as state sets is really a set, and in fact it
1619: is a subset of $(2^\omega )^3\times \cup _{Q\subseteq \omega
1620: }L^{Q\times \Sigma \times Q}.$ In most situations, however, the
1621: original version of Definition 3.2 is much more convenient and
1622: compatible with the corresponding definition in classical automata
1623: theory.
1624:
1625: Before investigating carefully various properties of regular
1626: $\ell-$valued languages, we present some interesting examples. The
1627: first one indicates that every $\ell-$valued language is regular.
1628: It is well-known that a similar conclusion holds in classical
1629: automata theory.
1630:
1631: \smallskip\
1632:
1633: \textbf{Example 3.1.} For any $A\in L^{\Sigma ^{*}},$ if $A$ is
1634: finite, i..e., $suppA=\{s\in \Sigma^{\ast}:A(s)>0\}$ is finite,
1635: then
1636: $$\stackrel{\ell }{\models }Reg_\Sigma (A).$$
1637: Indeed, suppose that $suppA=\{\sigma _{i1}...\sigma
1638: _{im_i}:i=1,...,k\}.$ Then we construct an $\ell -$valued
1639: automaton $\Re _A=(Q_A,I_A,T_A,\delta _A)$ in the following way:
1640:
1641: (i) $Q_A=\cup _{i=1}^k\{q_{i0},q_{i1},...,q_{im_i}\};$
1642:
1643: (ii) $I_A=\{q_{10},q_{20},...,q_{k0}\};$
1644:
1645: (iii) $T_A=\{q_{1m_1},q_{2m_2},...,q_{km_k}\};$ and
1646:
1647: (iv) We define $$\delta _A(q_{ij},\sigma
1648: _{i(j+1)},q_{i(j+1)})=A(\sigma _{i1}...\sigma _{im_i})$$ for any
1649: $1\leq i\leq k$ and $0\leq j<m_i,$ and we define $\delta
1650: _A(p,\sigma ,q)=0$ for other $(p,\sigma ,q)\in Q_A\times \Sigma
1651: \times Q_A.$ Then it is easy to see that $rec_{\Re _A}=A$ and
1652: $$\lceil Reg_\Sigma (A)\rceil \geq \lceil A\equiv rec_{\Re
1653: _A}\rceil =1.\heartsuit $$
1654:
1655: \smallskip\
1656:
1657: The following example may be seen as an extension of Example 3.1,
1658: and it shows that the recognizability of a quantum language is not
1659: less than the volume of its finite part.
1660:
1661: \smallskip\
1662:
1663: \textbf{Example 3.2.} For any $A\in L^{\Sigma ^{*}},$ we define
1664: $$A\downarrow \lambda =\{s\in \Sigma ^{*}:A(s)\not\leq \lambda
1665: \},$$ and
1666: $$A\uparrow \lambda =\{s\in \Sigma ^{*}:A(s)\not\geq
1667: \lambda \}.$$
1668:
1669: Let $A\in L^{\Sigma ^{*}}.$ Then
1670:
1671: (1) $\stackrel{\ell }{\models }\mathbf{\mu }\rightarrow Reg_\Sigma
1672: (A),$ where $\mu =\vee \{\lambda ^{\bot }:A\downarrow \lambda $ is
1673: finite$\};$ and
1674:
1675: (2) $\stackrel{\ell }{\models }\mathbf{\theta }\rightarrow
1676: Reg_\Sigma (A),$ where $\theta =\vee \{\lambda :A\uparrow \lambda
1677: $ is finite$\}.$
1678:
1679: Here, $\rightarrow$ may be interpreted as any implication operator
1680: satisfying the Birkhoff-von Neumann requirement. We only prove (1)
1681: and (2) may be proven likewise. For any $\lambda \in L,$ if
1682: $A\downarrow \lambda $ is finite, then we define $A\Downarrow
1683: \lambda \in L^{\Sigma ^{*}}$ as follows: for any $s\in \Sigma
1684: ^{*},$
1685: $$(A\Downarrow \lambda )(s)=\left\{
1686: \begin{array}{c}
1687: A(s)\ {\rm if}\ A(s)\not\leq \lambda , \\
1688: 0\ {\rm if}\ A(s)\leq \lambda .
1689: \end{array}
1690: \right. $$ Clearly, $A\Downarrow \lambda $ is finite. Then from
1691: Example 3.1 we know that there is an $\ell -$valued automata $\Re
1692: [\lambda ]$ such that $rec_{\Re [\lambda ]}=A\Downarrow \lambda ,$
1693: i.e., $ rec_{\Re [\lambda ]}=A(s)$ if $A(s)\not\leq \lambda $ and
1694: $rec_{\Re [\lambda ]}=0$ if $A(s)\leq \lambda ,$ and
1695: $$\lceil Rec_\Sigma (A)\rceil \geq \lceil A\equiv rec_{\Re [\lambda
1696: ]}\rceil =\wedge \{A(s)\leftrightarrow rec_{\Re [\lambda
1697: ]}:A(s)\not\leq \lambda \}\wedge \wedge \{A(s)\leftrightarrow
1698: 0:A(s)\leq \lambda \}$$
1699: $$=\wedge \{A(s)\leftrightarrow 0:A(s)\leq
1700: \lambda \}\geq \lambda ^{\bot }.\heartsuit $$
1701:
1702: \smallskip\
1703:
1704: The third example gives a simple connection between
1705: recognizability in classical automata theory and the $\ell
1706: -$valued predicate $Reg_\Sigma $ introduced above.
1707:
1708: \smallskip\
1709:
1710: \textbf{Example 3.3.} Let $A\subseteq \Sigma ^{*}$ be a regular
1711: language (in classical automata theory), $B\in L^{\Sigma ^{*}}$
1712: and
1713: $$suppB=\{s\in \Sigma ^{*}:B(s)>0\}\subseteq A,$$ and let
1714: $$\lambda =\vee \{\wedge _{s\in A}(a\leftrightarrow B(s)):a\in
1715: L\}.$$ Then $$\stackrel{\ell }{\models }\mathbf{\lambda
1716: }\rightarrow Reg_\Sigma (B).$$ In particular, if $A\subseteq
1717: \Sigma ^{*}$ is regular then for every $a \in L,$
1718: $$\stackrel{\ell }{\models }Reg_\Sigma (A[a]),$$ where $
1719: A[a]\in L^{\Sigma ^{*}}$ is given as
1720: $$A[a](s)=\left\{
1721: \begin{array}{c}
1722: a \ {\rm if}\ s\in A, \\
1723: 0\ {\rm otherwise.}
1724: \end{array}
1725: \right. $$
1726:
1727: This conclusion is not difficult to prove. In fact, since $A$ is
1728: regular, there must be an automaton $\Re =<Q,I,T,E>$ that accepts
1729: the language $A.$ Now, for each $a\in L,$ we construct an $ \ell
1730: -$valued automaton $\Re _a=<Q,I,T,\delta _a)$ such that
1731: $$\delta _a(p,\sigma ,q)=\left\{
1732: \begin{array}{c}
1733: a\ {\rm if}\ (p,\sigma ,q)\in E, \\
1734: 0\ {\rm otherwise.}
1735: \end{array}
1736: \right.$$ Then it is easy to know that for all $s\in \Sigma ^{*},$
1737: $$\lceil rec_{\Re _a}(s)\rceil =\left\{
1738: \begin{array}{c}
1739: a\ {\rm if}\ s\in A, \\
1740: 0\ {\rm otherwise,}
1741: \end{array}
1742: \right.$$ and
1743: $$\lceil B\equiv rec_{\Re _a}\rceil =\wedge _{s\in
1744: A}(a\leftrightarrow B(s)]).$$ Therefore, we have
1745: $$\lceil Reg_\Sigma
1746: (B)\rceil \geq \vee \{\lceil B\equiv rec_{\Re _a}\rceil :a\in
1747: L\}=\lambda .\heartsuit $$
1748:
1749: \smallskip\
1750:
1751: The fourth example demonstrates that the $\ell -$valued predicate
1752: $ Reg_\Sigma $ defined above is not trivial; that is, it does not
1753: in general degenerate into a two-valued (Boolean) predicate.
1754:
1755: \smallskip\
1756:
1757: \textbf{Example 3.4}. We consider a canonical orthomodular
1758: lattice. This lattice has a clear interpretation in quantum
1759: physics. One pasts together observables of the spin one-half
1760: system. Then he will obtain an orthomodular lattice $L(x)\oplus
1761: L(\overline{x}),$ where
1762: $$L(x)=\{0,p_{-},p_{+},1\}$$ corresponds to the outcomes of a
1763: measurement of the spin states along the $x-$axis and
1764: $$L(\overline{x})=\{
1765: \overline{0}=1,\overline{p_{-}},\overline{p_{+}},\overline{1}=0\}$$
1766: is obtained by measuring the spin states along a different spatial
1767: direction; and $L(x)\oplus L(\overline{x})$ may be visualized as
1768: the following ''Chinese lantern'' (see [Sv98] for a more detailed
1769: description of $ L(x)\oplus L(\overline{x})$) (see Figure 2).
1770:
1771: \begin{figure}\centering
1772: \includegraphics{fig_2.eps}\caption{"Chinese lantern"} \label{fig 2}
1773: \end{figure}
1774:
1775: In this example, we set $\rightarrow =\rightarrow_3$ (the
1776: Sasaki-hook). By a routine calculation we have
1777: $$p_{-}\leftrightarrow p_{+}=p_{-}\leftrightarrow
1778: \overline{p_{-}}=p_{-}\leftrightarrow \overline{ p_{+}}=0$$ and
1779: $p_{-}\leftrightarrow 1=p_{-}.$ Thus, for each $\lambda \in
1780: L(x)\oplus L(\overline{x}),$ $\lambda \not\leq p_{-}$ implies $
1781: p_{-}\leftrightarrow \lambda \leq p_{-}.$
1782:
1783: Furthermore, let $\Sigma =\{\sigma ,\tau \}$ and $A=\{\sigma
1784: ^n\tau ^n:n\in \omega \},$ and for any $t\in L(x)\oplus
1785: L(\overline{x}),$ let $A_t\in L^{\Sigma ^{*}}$ be given as
1786: follows:
1787: $$A_t(s)=\left\{
1788: \begin{array}{c}
1789: 1\ {\rm if}\ s\in A, \\
1790: t\ {\rm otherwise.}
1791: \end{array}
1792: \right. $$ Then it holds that $$\stackrel{\ell }{\models
1793: }\mathbf{p}_{-}\leftrightarrow Reg_\Sigma (A_{p_{-}});$$ that is,
1794: $\lceil Reg_\Sigma (A_{p_{-}})\rceil =p_{-}.$ In fact, we know
1795: that $\Sigma ^{*}$ is regular (see [E74], Example II.2.3), and
1796: with Example 3.3 it is easy to see that $\lceil Reg_\Sigma
1797: (A_{p_{-}})\rceil \geq p_{-}.$ Conversely, for any $\ell -$valued
1798: automaton $ \Re =<Q,I,T,\delta >,$ if $|Q|=n$ then
1799: $$\lceil A_{p_{-}}\equiv rec_\Re \rceil \leq [A_{p_{-}}(\sigma
1800: ^n\tau ^n)\leftrightarrow rec_\Re (\sigma ^n\tau ^n)]\wedge \wedge
1801: _{k,l\in \omega \ {\rm s.t. }\ k\neq l}[A_{p_{-}}(\sigma ^k\tau
1802: ^l)\leftrightarrow rec_\Re (\sigma ^k\tau ^l)]$$
1803: $$=rec_\Re (\sigma
1804: ^n\tau ^n)\wedge \wedge _{k,l\in \omega \ {\rm s.t. }\ k\neq
1805: l}[p_{-}\leftrightarrow rec_\Re (\sigma ^k\tau ^l)].$$ If $rec_\Re
1806: (\sigma ^n\tau ^n)\leq p_{-},$ then $\lceil A_{p_{-}}\equiv
1807: rec_\Re \rceil \leq p_{-}.$ Now, we consider the case of $rec_\Re
1808: (\sigma ^n\tau ^n)\not\leq p_{-}.$ For any $c\in T(Q,\Sigma ),$ if
1809: $b(c)\in I,$ $ e(c)\in T$ and $lb(c)=\sigma ^n\tau ^n,$ then $c$
1810: must be of the form $$ c=p_0\sigma p_1...p_{n-1}\sigma p_n\tau
1811: q_1...q_{n-1}\tau q_n.$$ Since $|Q|=n, $ there are $i,j$ such that
1812: $i<j\leq n$ and $p_i=p_j.$ We put
1813: $$c^{+}=p_0\sigma p_1...p_{j-1}\sigma p_j(=p_i)\sigma
1814: p_{i+1}...p_{j-1}\sigma p_j\sigma p_{j+1}...p_{n-1}\sigma p_n\tau
1815: q_1...q_{n-1}\tau q_n.$$ Then $b(c^{+})\in I,$ $e(c^{+})\in T,$
1816: $lb(c^{+})=\sigma ^{n+(j-i)}\tau ^n$ and $\lceil path_\Re
1817: (c^{+})\rceil =\lceil path_\Re (c)\rceil .$ Therefore, it holds
1818: that
1819: $$rec_\Re (\sigma ^{n+(j-i)}\tau ^n)\geq \vee \{\lceil
1820: path_\Re (c^{+})\rceil :b(c)\in I, e(c)\in T\ {\rm and}\
1821: lb(c)=\sigma ^n\tau ^n\}$$
1822: $$=\vee \{\lceil path_\Re (c)\rceil :b(c)\in I, e(c)\in T\ {\rm
1823: and}\ lb(c)=\sigma ^n\tau ^n\}$$
1824: $$=rec_\Re (\sigma ^n\tau ^n),$$
1825: and $$rec_\Re (\sigma ^{n+(j-i)}\tau ^n)\not\leq p_{-}.$$
1826: Furthermore, we have $$\lceil A_{p_{-}}\equiv rec_\Re \rceil \leq
1827: p_{-}\leftrightarrow rec_\Re (\sigma ^{n+(j-i)}\tau ^n)\leq
1828: p_{-}.$$
1829:
1830: So, for all $\ell -$valued automata $\Re $ we have $\lceil
1831: A_{p_{-}}\equiv rec_\Re \rceil \leq p_{-},$ and it follows that
1832: $$\lceil Rec_\Sigma (A_{p_{-}})\rceil =\vee \{\lceil A\equiv
1833: rec_\Re \rceil :\Re \in \mathbf{A} (\Sigma ,\ell )\}\leq p_{-}.$$
1834: This together with $\lceil Reg_\Sigma (A_{p_{-}})\rceil \geq
1835: p_{-}$ obtained before leads to $\lceil Reg_\Sigma
1836: (A_{p_{-}})\rceil =p_{-}.$
1837:
1838: Similarly, we have $\lceil Reg_\Sigma (A_t)\rceil =t$ for
1839: $t=p_{+},\overline{ p_{-}}$ and $\overline{p_{+}}.\heartsuit$
1840:
1841: \smallskip\
1842:
1843: Motivated by the above example, we propose the open problem: how
1844: to describe orthomodular lattices $\ell =<L,\leq ,\wedge ,\vee
1845: ,\bot ,0,1>$ which satisfy that
1846: $$\{\lceil Reg_\Sigma (A)\rceil :A\in L^{\Sigma ^{*}}\}=L,$$ i.e., the truth
1847: values of recognizability traverse all over $L,$ or more
1848: explicitly, for every $\lambda \in L,$ there is $A\in L^{\Sigma
1849: ^{*}}$ such that $\lceil Reg_\Sigma (A)\rceil =\lambda .\ $ It
1850: seems that this is a difficult problem.
1851:
1852: The $\ell-$valued regularity predicate $Reg_\Sigma$ in Definition
1853: 3.2 is a direct generalization of the notion of regular language
1854: in classical automata theory. In what follows, we will see that
1855: the predicate $Reg_\Sigma$ does not work well in many cases. Why
1856: this happens? Note that $Reg_\Sigma$ is merely a simple mimic of
1857: the classical concept of regular language, and an essential
1858: feature of quantum logic is missing here. In the defining equation
1859: of $Reg_\Sigma$, the language $A$ to be recognized and the
1860: automaton $\Re$ for recognizing $A$ are left completely
1861: irrelevant. In the case of classical logic, this does not causes
1862: any difficulty in manipulating regular languages. Nevertheless,
1863: the thing changes when we work in quantum logic. After an analysis
1864: it was found that a suitable link between $A$ and $\Re$ is a
1865: commutativity of them. This motivates the following:
1866:
1867: \smallskip\
1868:
1869: \textbf{Definition 3.3.} The $\ell -$valued (unary and partial)
1870: predicate $CReg_{\Sigma }$ on $L^{\Sigma ^{\ast }}$ is called
1871: commutative regularity and it is defined as $CReg_{\Sigma }\in
1872: L^{(L^{\Sigma ^{\ast }})}:$ for any $A\in L^{\Sigma ^{\ast }}$
1873: with finite $ Range(A)=\{A(s):s\in \Sigma ^{\ast }\},$
1874: $$CReg_{\Sigma }(A)\stackrel{def}{=}(\exists \Re
1875: \in \mathbf{A}(\Sigma ,\ell ))(\gamma (atom(\Re )\cup r(A))\wedge
1876: (A\equiv rec_{\Re })),$$ where $r(A)=\{\mathbf{a}:a\in
1877: Range(A)\}.$
1878:
1879: \smallskip\
1880:
1881: The exposition concerning the automata variable $\Re$ in the
1882: defining equation of $Reg_{\Sigma}$ in Definition 3.2 also applies
1883: to $CReg_{\Sigma}$ in the above definition.
1884:
1885: It is obvious that the notion of commutative regularity is
1886: stronger than regularity. In other words, we have for any $A\in
1887: L^{\Sigma^{\ast}}$,
1888: $$\stackrel{\ell }{\models } CReg_{\Sigma}(A)\rightarrow
1889: Reg_{\Sigma}(A).$$ On the other hand, if $\ell$ is a Boolean
1890: algebra; that is, the underlying logic is the classical Boolean
1891: logic, then these two notions are equivalent; or formally, for all
1892: $A\in L^{\Sigma^{\ast}}$, it holds that
1893: $$\stackrel{\ell }{\models } CReg_{\Sigma}(A)\leftrightarrow
1894: Reg_{\Sigma}(A).$$ This is just why the predicate $Reg_\Sigma$
1895: works very well in classical automata theory but not in the theory
1896: of automata based on quantum logic.
1897:
1898: \bigskip\
1899:
1900: \textbf{4. Orthomodular Lattice-Valued Deterministic Automata}
1901:
1902: \smallskip\
1903:
1904: The notion of nondeterminism plays a central role in the theory of
1905: computation. The nondeterministic mechanism enables a device to
1906: change its states in a way that is only partially determined by
1907: the current state and input symbol. Obviously, the concept of
1908: $\ell-$valued automaton introduced in the last section is a
1909: generalization of nondeterministic finite automaton. In classical
1910: theory of automata, each nondeterministic finite automaton is
1911: equivalent to a deterministic one; more exactly, there exists a
1912: deterministic finite automaton which accepts the same language as
1913: the originally given nondeterministic one does. The aim of this
1914: section is just to see whether this result is still valid in the
1915: framework of quantum logic. To this end, we first introduce the
1916: concept of deterministic $\ell-$valued automaton.
1917:
1918: Let $\Re=<Q,I,T,\delta>\in \mathbf{A}(\Sigma,\ell)$ be an
1919: $\ell-$valued automata over $\Sigma$. If
1920:
1921: (i) there is a unique $q_0$ in $Q$ with $I(q_0)>0$; and
1922:
1923: (ii) for all $q$ in $Q$ and $\sigma$ in $\Sigma ,$ there is a
1924: unique state $p$ in $Q$ such that $$ \delta (q,\sigma,p)>0,$$ then
1925: $M$ is called an $\ell-$valued (quantum) deterministic finite
1926: automaton ($\ell-$valued DFA for short). The $\ell-$valued
1927: transition relation $\delta $ in an $\ell-$valued DFA may be
1928: equivalently represented by a mapping from $Q\times \Sigma $ into
1929: $Q\times(L-\{0\}).$ For any $q$ in $Q$ and $\sigma$ in $\Sigma,$
1930: if $p$ is the unique element in $Q$ with $\delta (q,\sigma,p)>0,$
1931: then $\delta (q,\sigma)=(p,\delta (q,\sigma,p))\in
1932: Q\times(L-\{0\}).$
1933:
1934: The class of $\ell-$valued DFAs over $\Sigma$ is denoted
1935: $\mathbf{DA}(\Sigma,\ell)$.
1936:
1937: Suppose that $\Re$ is an $\ell-$valued DFA, $\delta
1938: (q_0,\sigma_1)=(q_1,\lambda _1)$ and $\delta
1939: (q_i,\sigma_{i+1})=(q_{i+1},\lambda _{i+1})$ for all
1940: $i=1,2,...,n-1.$ Then it is easy to see that $$\lceil rec_\Re
1941: (\sigma_1 ...\sigma_n)\rceil =I(q_0)\wedge T(q_n)\wedge \wedge
1942: _{i=1}^n\lambda _i.$$
1943:
1944: Throughout this section, we always suppose that the lattice $\ell$
1945: of truth values is finite.
1946:
1947: The proof of the equivalence between classical deterministic
1948: finite and nondeterministic finite automata is carried out by
1949: building the power set construction of a nondeterministic finite
1950: automaton that is deterministic and can simulate the given
1951: nondeterministic one. The power set construction can be naturally
1952: extended into the case of $\ell-$valued automata.
1953:
1954: Let $\Re =<Q,I,T,\delta
1955: >\in \mathbf{A}(\Sigma ,\ell )$ be an $\ell -$valued automaton
1956: over $\Sigma .$ We define the $\ell -$valued power set
1957: construction of $\Re $ to be $\ell -$valued automaton
1958: $$\ell ^{\Re }=<L^{Q},I_{1},\overline{T},\overline{
1959: \delta }>$$ over $\Sigma ,$ where:
1960:
1961: (i) $L^{Q}$ is the set of all $\ell -$valued subsets of $Q;$ that
1962: is, mappings from $Q$ into $L;$
1963:
1964: (ii) $I_{1}$ is an $\ell -$valued point with height $1;$ that is,
1965: $I_{1}\in L^{(L^{Q})}$ and
1966: $$I_{1}(X)=\left\{
1967: \begin{array}{c}
1968: 1\ {\rm if }X=I, \\
1969: 0\ {\rm otherwise}
1970: \end{array}
1971: \right.$$ for all $X\in L^{Q};$
1972:
1973: (iii) $\overline{T}\in L^{(L^{Q})};$ that is, $\overline{T}$ is an
1974: $\ell -$valued subset of $L^{Q},$ and
1975: $$\overline{T}(X)=\vee _{q\in Q}[X(q)\wedge T(q)]$$
1976: for any $X\in L^{Q};$ and
1977:
1978: (iv) $\overline{\delta }$ is a mapping from $L^{Q}\times \Sigma $
1979: into $ L^{Q},$ and for each $X\in L^{Q},$ $\overline{\delta
1980: }(X,\sigma )\in L^{Q}$ and
1981: $$\overline{\delta }(X,\sigma )(q)=\vee _{p\in Q}[X(p)\wedge \delta (p,\sigma
1982: ,q)]$$ for every $q\in Q.$
1983:
1984: Since $L$ is assumed to be finite, $L^{Q}$ is finite too. Thus, it
1985: is easy to see that $\ell ^{\Re }$ is an $\ell -$valued DFA.
1986: Moreover, both the set of the initial states and the transition
1987: relation are two-valued, namely, their truth values are either $0$
1988: or $1,$ and only the set of terminal states carries $\ell -$valued
1989: information.
1990:
1991: The following theorem compares the abilities of an $\ell-$valued
1992: automaton and its power set construction according to the
1993: $\ell-$valued languages recognized by them.
1994:
1995: \smallskip\
1996:
1997: \textbf{Theorem 4.1.} Let $\ell =<L,\leq ,\wedge ,\vee ,\bot
1998: ,0,1>$ be a finite orthomodular lattice, and let $\longrightarrow$
1999: be an implication operator satisfying the Birkhoff-von Neumann
2000: requirement.
2001:
2002: (1) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and $s\in \Sigma
2003: ^{\ast },$
2004: $$\stackrel{\ell }{ \models } rec_{\Re }(s)\longrightarrow rec_{(\ell
2005: ^{\Re })}(s).$$
2006:
2007: (2) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and $s\in \Sigma
2008: ^{\ast },$
2009: $$\stackrel{\ell }{ \models } \gamma(atom(\Re))\wedge rec_{(\ell
2010: ^{\Re })}(s) \longrightarrow rec_{\Re }(s),$$ and in particular if
2011: $\rightarrow =\rightarrow_3$, then
2012: $$\stackrel{\ell }{ \models } \gamma(atom(\Re))\longrightarrow (rec_{(\ell
2013: ^{\Re })}(s) \longleftrightarrow rec_{\Re }(s)).$$
2014:
2015: (3) The following two statements are equivalent to each other:
2016:
2017: \ \ \ \ \ \ (i) $\ell$ is a Boolean algebra.
2018:
2019: \ \ \ \ \ \ (ii) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and
2020: $s\in \Sigma ^{\ast },$
2021: $$\stackrel{\ell }{ \models } rec_{\Re }(s)\longleftrightarrow rec_{(\ell
2022: ^{\Re })}(s).$$
2023:
2024: \smallskip\
2025:
2026: \textbf{Proof.} The proof of (1) is easy, and we omit it here.
2027:
2028: (2) Suppose that $\Re =<Q,I,T,\delta
2029: >\in \mathbf{A}(\Sigma ,\ell )$ and $\ell ^{\Re
2030: }=<L^{Q},I_{1},\overline{T}, \overline{\delta }>$ is the $\ell
2031: -$valued power set construction of $\Re .$ Our aim is to
2032: demonstrate that
2033: $$\lceil \gamma(atom(\Re))\rceil \wedge \lceil rec_{(\ell ^{\Re })}(s)\rceil\leq \lceil rec_{\Re }(s)\rceil
2034: $$ for all $s\in \Sigma ^{\ast }.$ To this end, we first prove the
2035: following
2036: $$claim:\ \ \lceil \gamma(atom(\Re))\rceil \wedge \overline{\delta }(I,\sigma _{1}...\sigma
2037: _{n})(q_{n}) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
2038: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
2039: $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \leq\vee \{I(q_{0})\wedge \wedge
2040: _{i=0}^{n-1}\delta (q_{i},\sigma
2041: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{n-1}\in Q\}$$ for any $\sigma
2042: _{1},...,\sigma _{n}\in \Sigma $ and $q\in Q.$ We proceed by
2043: induction on $n.$ For $n=0,$ it is clear. The definition of
2044: $\overline{\delta}$ yields
2045: $$\overline{\delta }(I,\sigma _{1}...\sigma
2046: _{n})(q_{n})=\overline{\delta }( \overline{\delta }(I,\sigma
2047: _{1}...\sigma _{n-1}),\sigma _{n})(q_{n})\ \ \ \ \ \ \ \ \ \ \ \ \
2048: \ \ \ \ \ \ \ \ \ $$
2049: $$=\vee _{q_{n-1}\in Q}[\overline{\delta }(I,\sigma _{1}...\sigma
2050: _{n-1})(q_{n-1})\wedge \delta (q_{n-1},\sigma _{n},q_{n})].$$ We
2051: write $$\lceil atom(\Re)\rceil =\{\lceil \varphi\rceil: \varphi\in
2052: atom(\Re)\}.$$ Then it holds that
2053: $$\lceil \gamma(atom(\Re))\rceil=\gamma(\lceil atom(\Re)\rceil).$$ Note
2054: that the symbol $\gamma$ in the left-hand side applies to a set of
2055: logical formulas, whereas the one in the right-hand side applies
2056: to a subset of $L$. Furthermore, it is easy to see that
2057: $\delta(q_{n-1},\sigma_n,q_n),\ \overline{\delta}(I,\sigma_1
2058: ...\sigma_{n-1})(q_{n-1})$ and $\lceil \gamma(atom(\Re))\rceil$
2059: are all in $[\lceil atom(\Re)\rceil]$ (the subalgebra of $\ell$
2060: generated by $\lceil atom(\Re)\rceil$). Thus, with Lemmas 2.5 and
2061: 2.6 and the induction hypothesis we obtain
2062: $$\lceil \gamma(atom(\Re))\rceil\wedge \overline{\delta }(I,\sigma _{1}...\sigma
2063: _{n})(q_{n}) =\lceil \gamma(atom(\Re))\rceil\wedge \lceil
2064: \gamma(atom(\Re))\rceil $$ $$\ \ \ \ \ \ \ \ \ \ \wedge \vee
2065: _{q_{n-1}\in Q}[\overline{\delta }(I,\sigma _{1}...\sigma
2066: _{n-1})(q_{n-1})\wedge \delta (q_{n-1},\sigma _{n},q_{n})]$$
2067: $$\leq \vee _{q_{n-1}\in Q}[\lceil \gamma(atom(\Re))\rceil\wedge
2068: (\vee \{I(q_0)\wedge\wedge_{i=0}^{n-2}\delta(q_i,\sigma
2069: _{i+1},q_{i+1}):$$ $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
2070: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
2071: q_0,q_1,...,q_{n-2}\in Q\})\wedge \delta (q_{n-1},\sigma
2072: _{n},q_{n})].$$ Using Lemmas 2.5 and 2.6 again, we complete the
2073: proof of the above claim.
2074:
2075: Now with this claim, we can use Lemmas 2.5 and 2.6 twice and
2076: derive that
2077: $$\lceil \gamma(atom(\Re))\rceil\wedge \lceil
2078: rec_{(\ell^{Re})}(\sigma_1 ...\sigma_n)\rceil =\lceil
2079: \gamma(atom(\Re))\rceil\wedge \overline{T}( \overline{\delta
2080: }(I,\sigma _{1}...\sigma _{n}))$$
2081: $$=\lceil \gamma(atom(\Re))\rceil\wedge \vee _{q_{n}\in Q}[\overline{\delta }(I,\sigma _{1}...\sigma
2082: _{n})(q_{n})\wedge T(q_{n})]$$
2083: $$\leq \vee _{q_{n}\in Q}[\lceil \gamma(atom(\Re))\rceil\wedge \overline{\delta }(I,\sigma _{1}...\sigma
2084: _{n})(q_{n})\wedge T(q_{n})]$$
2085: $$\leq \vee _{q_{n}\in Q}[\lceil \gamma(atom(\Re))\rceil\wedge (\vee \{I(q_{0})\wedge \wedge
2086: _{i=0}^{n-1}\delta (q_{i},\sigma
2087: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{n-1}\in Q\})\wedge T(q_{n})]$$
2088: $$\leq \vee \{I(q_{0})\wedge \wedge _{i=0}^{n-1}\delta (q_{i},\sigma
2089: _{i+1},q_{i+1})\wedge T(q_{n}):q_{0},q_{1},...,q_{n-1}\in Q\}$$
2090: $$=\lceil rec_{\Re }(\sigma _{1}...\sigma _{n})\rceil .$$
2091:
2092: For the case of $\rightarrow =\rightarrow_3$, what we want to
2093: prove is
2094: $$\lceil \gamma(atom(\Re))\rceil \leq \lceil rec_{(\ell ^{\Re })}(s)\rceil\longrightarrow_3 \lceil rec_{\Re }(s)\rceil
2095: .$$ With the above conclusion and Lemma 2.10, it suffices to show
2096: that $\lceil \gamma(atom(\Re))\rceil C \lceil rec_{\Re
2097: }(s)\rceil$. We observe that $$\lceil \gamma(atom(\Re))\rceil
2098: =\vee \{\wedge_{\varphi\in \lceil
2099: atom(\Re)\rceil}\varphi^{f(\varphi)}:f\in \{0,1\}^{\lceil
2100: atom(\Re)\rceil}\}.$$ Then Lemma 2.2 tells us that we only need to
2101: prove $\wedge_{\varphi\in \lceil
2102: atom(\Re)\rceil}\varphi^{f(\varphi)}C \lceil rec_{\Re }(s)\rceil$
2103: for all $f\in \{0,1\}^{\lceil atom(\Re)\rceil}$. For every
2104: $\psi\in \lceil atom(\Re)\rceil$, note that
2105: $$\wedge_{\varphi\in \lceil
2106: atom(\Re)\rceil}\varphi^{f(\varphi)}\leq \psi^{f(\psi)}.$$ Then
2107: $\wedge_{\varphi\in \lceil atom(\Re)\rceil}\varphi^{f(\varphi)} C
2108: \psi^{f(\psi)},$ and furthermore it follows that
2109: $\wedge_{\varphi\in \lceil atom(\Re)\rceil}\varphi^{f(\varphi)} C
2110: \psi$ from Lemmas 2.1(3) and (4). Since $\lceil rec_\Re (s)\rceil$
2111: is calculated from some elements in $\lceil atom(\Re)\rceil$ by
2112: applying a finite number of meets or unions, we complete the proof
2113: with Lemma 2.2.
2114:
2115: (3) Note that $\lceil \gamma(atom(\Re))\rceil=1$ is always valid
2116: when $\ell$ is a Boolean algebra. Thus, it is proved that (i)
2117: implies (ii). We now turn to show that (ii) implies (i). It
2118: suffices to show that the meet $\wedge$ is distributive over the
2119: union $\vee$; that is, $$a\wedge (b\vee c)=(a\wedge b)\vee
2120: (a\wedge c)$$ for all $a,b,c\in L$. Let $a,b,c\in L.$ We construct
2121: an $\ell -$valued automaton
2122: $$\Re =<\{u,v,w\},\{u,v\},\{w\},\delta >$$ over $ \Sigma $ which
2123: has at least one element $\sigma ,$ where $\delta (u,\sigma
2124: ,u)=a,$ $\delta (u,\sigma ,w)=c,$ $\delta (v,\sigma ,u)=b,$ and
2125: $\delta $ takes the value $0$ for other cases. It may be
2126: visualized by Figure 3.
2127:
2128: \begin{figure}\centering
2129: \includegraphics[width=0.5\textwidth]{fig_3.eps}\caption{Automaton a} \label{fig 3}
2130: \end{figure}
2131:
2132: In the automaton $\Re $ we have
2133: $$\lceil rec_{\Re }(\sigma \sigma )\rceil =\vee \{I(q_{0})\wedge
2134: T(q_{2})\wedge \delta (q_{0},\sigma ,q_{1})\wedge \delta
2135: (q_{1},\sigma ,q_{2}):q_{0},q_{1},q_{2}\in Q\}$$
2136: $$=\vee \{\delta (u,\sigma ,q_{1})\wedge \delta (q_{1},\sigma
2137: ,w):q_{1}\in Q\}\vee \vee \{\delta (v,\sigma ,q_{1})\wedge \delta
2138: (q_{1},\sigma ,v):q_{1}\in Q\}$$
2139: $$=[\delta (u,\sigma ,u)\wedge \delta (u,\sigma ,w)]\vee \lbrack
2140: \delta (v,\sigma ,u)\wedge \delta (u,\sigma ,w)]$$
2141: $$=(a\wedge c)\vee (b\wedge c).$$
2142:
2143: Consider the $\ell -$valued power set construction $\ell ^{\Re }$
2144: of $\Re .$ Then
2145: $$\overline{\delta }(I,\sigma )(u)=\vee _{q\in Q}[I(q)\wedge \delta
2146: (q,\sigma ,u)]$$
2147: $$=\delta (u,\sigma ,u)\vee \delta (v,\sigma ,u)]$$
2148: $$=a\vee b.$$
2149: Similarly, we obtain $\overline{\delta }(I,\sigma )(v)=0$ and
2150: $\overline{ \delta }(I,\sigma )(w)=c.$ It follows that for any
2151: $q\in Q,$
2152: $$\overline{\delta }(I,\sigma \sigma )(q)=\overline{\delta
2153: }(\overline{\delta }(I,\sigma ),\sigma )(q)$$
2154: $$=\vee _{q^{\prime }\in Q}[\overline{\delta }(I,\sigma )(q^{\prime
2155: })\wedge \delta (q^{\prime },\sigma ,q)]$$
2156: $$=[\overline{\delta }(I,\sigma )(u)\wedge \delta (u,\sigma
2157: ,q)]\vee \lbrack \overline{\delta }(I,\sigma )(w)\wedge \delta
2158: (w,\sigma ,q)]$$
2159: $$=(a\vee b)\wedge \delta (u,\sigma ,q).$$
2160: Thus, $$\overline{\delta }(I,\sigma \sigma )(u)=(a\vee b)\wedge
2161: a=a,$$ $$ \overline{\delta }(I,\sigma \sigma )(v)=0$$ and
2162: $$\overline{\delta }(I,\sigma \sigma )(w)=(a\vee b)\wedge c.$$
2163: Therefore,
2164: $$\lceil rec_{(\ell ^{\Re })}(\sigma \sigma )\rceil
2165: =\overline{T}(\overline{ \delta }(I,\sigma \sigma ))$$
2166: $$=\vee _{q\in Q}[\overline{\delta }(I,\sigma \sigma )(q)\wedge
2167: T(q)]$$
2168: $$=\overline{\delta }(I,\sigma \sigma )(w)$$
2169: $$=(a\vee b)\wedge c.$$
2170:
2171: Finally, from the assumption (ii) we assert that
2172: $$(a\wedge c)\vee (b\wedge c)=\lceil rec_{\Re }(\sigma \sigma
2173: )\rceil =\lceil rec_{(\ell ^{\Re })}(\sigma \sigma )\rceil =(a\vee
2174: b)\wedge c.\heartsuit$$
2175:
2176: \smallskip\
2177:
2178: Many results in this paper appear in the same scheme as the above
2179: theorem. So, we here give a detailed explanation of this theorem.
2180: The above theorem points out that the ability of an $\ell-$valued
2181: automaton for recognizing language is always weaker than that of
2182: its power set construction. On the other hand, in order to warrant
2183: that an $\ell-$valued automaton $\Re$ and its power set
2184: construction have the same ability of accepting language, the
2185: condition $\gamma(atom(\Re))$ has to be imposed. The intuitive
2186: meaning of this condition is that (the truth values of) any two
2187: atomic propositions describing $\Re$ should commute. (See also the
2188: physical interpretation of commutativity presented in the
2189: introductory section.) The third part of Theorem 4.1 indicates
2190: that the equivalence between a nondeterministic finite automaton
2191: and its power set construction is universally valid if and only if
2192: the underlying logic degenerates to the classical Boolean logic.
2193: In other words, if the meta-logic that we use in our reasoning
2194: does not enjoy distributivity, then such a meta-logic is not
2195: strong enough to guarantee the universal validity of any
2196: nondeterministic finite automaton and its power set construction,
2197: and we can always find a nondeterministic finite automaton such
2198: that the equivalence between it and its power set construction is
2199: not derivable with the mere inference power provided by such a
2200: meta-logic.
2201:
2202: In Section 3, we introduced the regularity and commutative
2203: regularity predicates $Reg_{\Sigma}$ and $CReg_{\Sigma}$. They are
2204: all given with respect to nondeterministic $\ell-$valued automata.
2205: Now we propose a restricted version of them based on the smaller
2206: class of deterministic $\ell-$valued automata.
2207:
2208: \smallskip\
2209:
2210: \textbf{Definition 4.1.} Let $\ell=<L,\leq ,\wedge ,\vee ,\bot
2211: ,0,1>$ be an orthomodular lattice. Then the $\ell -$valued (unary)
2212: predicates $DReg_{\Sigma}$ and (unary and partial) predicate
2213: $CDReg_{\Sigma }$ on $L^{\Sigma ^{\ast }}$ are called
2214: deterministic regularity and commutative deterministic regularity,
2215: respectively, and they are defined as $DReg_{\Sigma}$,
2216: $CDReg_{\Sigma }\in L^{(L^{\Sigma ^{\ast }})}:$ for any $A\in
2217: L^{\Sigma ^{\ast }}$,
2218: $$DReg_{\Sigma }(A)\stackrel{def}{=}(\exists \Re
2219: \in \mathbf{DA}(\Sigma ,\ell ))(A\equiv rec_{\Re }),$$ and for any
2220: $A\in L^{\Sigma ^{\ast }}$ with finite $Range(A)=\{A(s):s\in
2221: \Sigma^{\ast}\},$
2222: $$CDReg_{\Sigma }(A)\stackrel{def}{=}(\exists \Re
2223: \in \mathbf{DA}(\Sigma ,\ell ))(\gamma (atom(\Re )\cup r(A))\wedge
2224: (A\equiv rec_{\Re })),$$ where $r(A)=\{\mathbf{a}:a\in
2225: Range(A)\}.$
2226:
2227: \smallskip\
2228:
2229: It is similar to the relation between $Reg_\Sigma$ and
2230: $CReg_\Sigma$ that $CDReg_\Sigma$ is stronger than $DReg_\Sigma$.
2231: In other words, it holds that for any $A\in L^{\Sigma^{\ast}}$,
2232: $$\stackrel{\ell }{\models } CDReg_{\Sigma}(A)\rightarrow
2233: DReg_{\Sigma}(A).$$ The following corollary shows that a certain
2234: commutativity condition guarantees that they are equivalent.
2235: Furthermore, if $\ell$ is a Boolean algebra, then the four notions
2236: $Reg_{\Sigma}$, $CReg_{\Sigma}$, $DReg_{\Sigma}$ and
2237: $CDReg_{\Sigma}$ all coincide.
2238:
2239: \smallskip\
2240:
2241: \textbf{Corollary 4.2.} Let $\ell=<L,\leq ,\wedge ,\vee ,\bot
2242: ,0,1>$ be a finite orthomodular lattice, and let $\longrightarrow
2243: = \longrightarrow_3$. Then for any $A\in L^{\Sigma^{\ast}}$,
2244: $$\stackrel{\ell }{ \models }CReg_{\Sigma}(A)\longleftrightarrow CDReg_{\Sigma}(A).$$
2245: In particular, if $\ell$ is a Boolean algebra, then for any $A\in
2246: L^{\Sigma ^{\ast }},$
2247: $$\stackrel{\ell }{ \models } Reg_{\Sigma}(A)\longleftrightarrow DReg_{\Sigma}(A) .$$
2248:
2249: \smallskip\
2250:
2251: \textbf{Proof}. It is clear that $$\stackrel{\ell }{ \models
2252: }CDReg_{\Sigma}(A)\rightarrow CDReg_{\Sigma }(A).$$ Then we only
2253: need to prove that $$\stackrel{\ell }{ \models }CReg_{\Sigma
2254: }(A)\rightarrow CDReg_{\Sigma }(A);$$ that is, for any $\Re \in
2255: \mathbf{A}(\Sigma ,\ell ),$
2256: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv
2257: rec_{\Re }\rceil \leq \vee \{\lceil \gamma (atom(\wp )\cup
2258: r(A))\rceil \wedge \lceil A\equiv rec_{\wp }\rceil :\wp \in
2259: \mathbf{DA}(\Sigma ,\ell )\}.$$
2260:
2261: First, by using Lemmas 2.5, 2.6 and 2.11(2) we have $$\lceil
2262: \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv rec_{\Re
2263: }\rceil \wedge \lceil rec_{\Re }\equiv rec_{(\ell ^{\Re })}\rceil
2264: =\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
2265: \ \ \ $$
2266: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \wedge _{s\in
2267: \Sigma ^{\ast }}(A(s)\leftrightarrow rec_{\Re }(s))\wedge \wedge
2268: _{s\in \Sigma ^{\ast }}(rec_{\Re }(s)\leftrightarrow rec_{(\ell
2269: ^{\Re })}(s))$$
2270: $$=\wedge _{s\in \Sigma ^{\ast }}(\lceil \gamma (atom(\Re )\cup
2271: r(A))\rceil \wedge (A(s)\rightarrow rec_{\Re }(s))\wedge (rec_{\Re
2272: }(s)\rightarrow rec_{(\ell ^{\Re })}(s))\wedge$$
2273: $$\wedge _{s\in \Sigma ^{\ast }}(\lceil \gamma (atom(\Re
2274: )\cup r(A))\rceil \wedge (rec_{(\ell ^{\Re })}(s)\rightarrow
2275: rec_{\Re }(s))\wedge (rec_{\Re }(s)\rightarrow A(s))$$
2276: $$\leq \wedge _{s\in \Sigma ^{\ast }}(A(s)\rightarrow rec_{(\ell
2277: ^{\Re })}(s))\wedge \wedge _{s\in \Sigma ^{\ast }}(rec_{(\ell
2278: ^{\Re })}(s)\rightarrow A(s))$$
2279: $$=\wedge _{s\in \Sigma ^{\ast }}(A(s)\leftrightarrow rec_{(\ell
2280: ^{\Re })}(s)) $$
2281: $$=\lceil A\equiv rec_{(\ell ^{\Re })}\rceil .$$
2282:
2283: Second, from Theorem 4.1(2) we obtain
2284: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \leq \lceil \gamma
2285: (atom(\Re ))\rceil \leq \lceil rec_{\Re }\equiv rec_{(\ell ^{\Re
2286: })}\rceil ,$$ and
2287: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv
2288: rec_{\Re }\rceil \leq \lceil \gamma (atom(\Re )\cup r(A))\rceil
2289: \wedge \lceil A\equiv rec_{\Re }\rceil \wedge \lceil rec_{\Re
2290: }\equiv rec_{(\ell ^{\Re })}\rceil $$
2291: $$\leq \lceil A\equiv rec_{(\ell ^{\Re })}\rceil .$$
2292:
2293: In addition, it is easy to see that
2294: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \leq \lceil \gamma
2295: (atom(\ell ^{\Re })\cup r(A))\rceil $$ from Lemma 2.6. Therefore,
2296: it follows that
2297: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv
2298: rec_{\Re }\rceil \leq \lceil \gamma (atom(\ell ^{\Re })\cup
2299: r(A))\rceil \wedge \lceil A\equiv rec_{(\ell ^{\Re })}\rceil $$
2300: $$\leq \vee \{\lceil \gamma (atom(\wp )\cup r(A))\rceil \wedge
2301: \lceil A\equiv rec_{\wp }\rceil :\wp \in \mathbf{DA}(\Sigma ,\ell
2302: )\},$$ and we complete the proof. $\heartsuit $
2303:
2304: \smallskip\
2305:
2306: It should be note that in the above corollary the second
2307: conclusion is obtained from the first one by removing simply the
2308: commutativity. The second conclusion is in general not correct.
2309: The reason is that an essential application is needed in the
2310: derivation of the implication $CReg_\Sigma\longrightarrow
2311: CDReg_\Sigma.$
2312:
2313: \bigskip\
2314:
2315: \textbf{5. Orthomodular Lattice-Valued Automata with
2316: $\varepsilon-$Moves}
2317:
2318: \smallskip\
2319:
2320: Automata with $\varepsilon-$moves are nondeterministic automata in
2321: which transitions on the empty input $\varepsilon$ are included,
2322: and they have the same power for accepting languages. In the
2323: classical theory of automata, automata with $\varepsilon-$moves
2324: are very convenient tools in building complex automata from simple
2325: ones and in proving the closure properties of regular languages.
2326: The aim of this section is to introduce an orthomodular
2327: lattice-valued extension of automaton with $\varepsilon-$moves.
2328: Let $\ell =<L,\leq ,\wedge ,\vee ,\perp ,0,1>$ be an orthomodular
2329: lattice. Then an $\ell -$valued automaton with $\varepsilon
2330: -$moves over $\Sigma$ is a quadruple $ \Re =<Q,I,T,\delta >$ in
2331: which all components are the same as in an $\ell -$valued
2332: automaton (without $\varepsilon -$moves), but the domain of the
2333: quantum transition relation $\delta $ is changed to $Q\times
2334: (\Sigma \cup \{\varepsilon \})\times Q;$ that is, $\delta $ is a
2335: mapping from $Q\times (\Sigma \cup \{\varepsilon \})\times Q$ into
2336: $L,$ where $\varepsilon $ stands for the empty string of input
2337: symbols. Thus, in an $\ell -$valued automaton with $\varepsilon
2338: -$moves, transitions of the form "$p\stackrel{\delta ,\varepsilon
2339: }{\longrightarrow }q$" are allowed. So, $atom(\Re )$ contains the
2340: atomic propositions "$p\stackrel{\delta ,\varepsilon
2341: }{\longrightarrow }q$", and their truth values are given as
2342: $\delta (p,\varepsilon ,q)$ for all $p,q\in Q.$
2343:
2344: Now let $\Re =<Q,I,T,\delta >$ be an $\ell -$valued automaton with
2345: $ \varepsilon -$moves. We put $$T_{\varepsilon }(Q,\Sigma
2346: )=(Q(\Sigma \cup \{\varepsilon \}))^{\ast }Q=\cup _{n=0}^{\infty
2347: \lbrack }(Q(\Sigma \cup \{\varepsilon \}))^{n}Q].$$ The difference
2348: between $T(Q,\Sigma )$ and $T_{\varepsilon }(Q,\Sigma )$ is that
2349: in the latter the empty string may be used as labels. For any $
2350: c=q_{0}\sigma_{1} q_{1}...q_{k-1}\sigma _{k}q_{k}\in
2351: T_{\varepsilon }(Q,\Sigma ), $ $lb_{\varepsilon }(c)$ is defined
2352: to be the sequence $\sigma _{1}...\sigma _{k} $ with all
2353: occurrences of $\varepsilon $ deleted. Note that it is possible
2354: that the length of $lb_{\varepsilon }(c)$ is strictly smaller than
2355: $k.$ Then the recognizability $ rec_{\Re }$ is also defined as an
2356: $\ell -$valued unary predicate over $ \Sigma ^{\ast },$ and it is
2357: given by
2358: $$rec_{\Re }(s)\stackrel{def}{=}(\exists c\in T_{\varepsilon
2359: }(Q,\Sigma ))(b(c)\in I\wedge e(c)\in T\wedge lb_{\varepsilon
2360: }(c)=s\wedge path_{\Re }(c))$$ for all $s\in \Sigma ^{\ast },$
2361: where $path_{\Re }$ is defined in the same way as in an $\ell
2362: -$valued automaton without $\varepsilon -$moves. The defining
2363: equation of $rec_\Re$ may be rewritten in terms of truth valued as
2364: follows: $$\lceil rec_\Re (s)\rceil = \vee \{I(b(c))\wedge
2365: T(e(c))\lceil path_\Re (c)\rceil : c\in T_\varepsilon (Q,\Sigma)\
2366: {\rm and}\ lb_\varepsilon (c)=s\},$$ where $$\lceil path_\Re
2367: (c)\rceil = \wedge_{i=0}^{k-1}\delta(q_i,\sigma_{i+1},q_{i+1})$$
2368: if $c=q_0 \sigma_1 q_1 ... q_{k-1} \sigma_k q_k.$
2369:
2370: For any $\ell -$valued automaton $\Re =<Q,I,T,\delta >$ with
2371: $\varepsilon -$ moves, its $\varepsilon-$reduction is defined to
2372: be the $\ell -$valued automaton $\Re ^{-\varepsilon
2373: }=<Q,I,T^{\prime },\delta ^{\prime }>$ (without $\varepsilon -
2374: $moves) in which
2375:
2376: (i) for any $q\in Q,$
2377: $$q\in T^{\prime }\stackrel{def}{=}(\exists q\in Q,m\geq 0)(q\in
2378: T\wedge \delta (q_{0},\varepsilon ^{m},q));$$ that is,
2379: $$T^{\prime }(q)=\vee _{q\in Q,m\geq 0}(T(q)\wedge \delta
2380: (q_{0},\varepsilon ^{m},q));$$
2381:
2382: (ii) for any $p,q\in Q$ and $\sigma \in \Sigma ,$ $$\delta
2383: ^{\prime }(p,\sigma ,q)\stackrel{def}{=}(\exists m,n\geq 0)\delta
2384: (p,\varepsilon ^{m}\sigma \varepsilon ^{n},q);$$ that is,
2385: $$\delta ^{\prime }(p,\sigma ,q)=\vee _{m,n\geq 0}\delta
2386: (p,\varepsilon ^{m}\sigma \varepsilon ^{n},q),$$ where
2387: $$\delta (q_{0},\sigma _{1}...\sigma
2388: _{k},q_{k})\stackrel{def}{=}(\exists q_{1},...,q_{k-1}\in
2389: Q)(\delta (q_{0},\sigma _{1},q_{1})\wedge \delta (q_{1},\sigma
2390: _{2},q_{2})\wedge \delta (q_{k-1},\sigma _{k},q_{k})).$$ In other
2391: words,
2392: $$\delta (q_{0},\sigma _{1}...\sigma _{k},q_{k})=\vee \{(\delta
2393: (q_{0},\sigma _{1},q_{1})\wedge \delta (q_{1},\sigma
2394: _{2},q_{2})\wedge \delta (q_{k-1},\sigma
2395: _{k},q_{k}):q_{1},...,q_{k-1}\in Q\}$$ for all $k\geq 1,$
2396: $q_{0},q_{k}\in Q$ and $\sigma _{1},...,\sigma _{k}\in \Sigma .$
2397:
2398: The following theorem gives a clear relation between the language
2399: accepted by an orthomodular lattice-valued automaton with
2400: $\varepsilon-$moves and that accepted by its
2401: $\varepsilon-$reduction. In general, the $\varepsilon-$reduction
2402: of an automaton with $\varepsilon-$moves has a stronger power of
2403: acceptance than itself. A certain commutativity between basic
2404: actions of the automaton implies the equivalence between an
2405: automaton with $\varepsilon-$moves and its
2406: $\varepsilon-$reduction. However, an universal validity of such an
2407: equivalence requires that the underlying logic degenerates to the
2408: classical Boolean logic.
2409:
2410: \smallskip\
2411:
2412: \textbf{Theorem 5.1.} Let $\ell =<L,\leq ,\wedge ,\vee ,\bot
2413: ,0,1>$ be an orthomodular lattice, and let $\longrightarrow$ be an
2414: implication operator satisfying the Birkhoff-von Neumann
2415: requirement.
2416:
2417: (1) For any $\ell -$valued automaton $\Re $ with $\varepsilon
2418: -$moves over $ \Sigma ,$ and for any $s\in \Sigma ^{\ast },$
2419: $$\stackrel{\ell }{\models }rec_{\Re }(s)\rightarrow rec_{\Re
2420: ^{-\varepsilon }}(s).$$
2421:
2422: (2) For any $\ell -$valued automaton $\Re $ with $\varepsilon
2423: -$moves over $ \Sigma ,$ and for any $s\in \Sigma ^{\ast },$
2424: $$\stackrel{\ell }{\models }\gamma (atom(\Re ))\wedge rec_{\Re
2425: ^{-\varepsilon }}(s)\rightarrow rec_{\Re }(s),$$ and in particular
2426: if $\rightarrow =\rightarrow _{3}$ then
2427: $$\stackrel{\ell }{\models }\gamma (atom(\Re ))\rightarrow (rec_{\Re
2428: }(s)\leftrightarrow rec_{\Re ^{-\varepsilon }}(s)).$$
2429:
2430: (3) The following two statements are equivalent:
2431:
2432: \ \ \ \ \ \ (i) $\ell $ is a Boolean algebra;
2433:
2434: \ \ \ \ \ \ (ii) For all $\ell -$valued automaton $\Re $ with
2435: $\varepsilon -$moves over $ \Sigma ,$ and for all $s\in \Sigma
2436: ^{\ast },$
2437: $$\stackrel{\ell }{\models }rec_{\Re }(s)\leftrightarrow rec_{\Re
2438: ^{-\varepsilon }}(s).$$
2439:
2440: \smallskip\
2441:
2442: \textbf{Proof.} The proof of (1) is similar to that of (2), so we
2443: omit it. We now prove (2). First, we use induction on the length
2444: $|c|$ of $c$ to show that for any $c\in T(Q,\Sigma ),$
2445: $$claim:\ \ \lceil \gamma (atom(\Re )\rceil \wedge \lceil path_{\Re
2446: ^{-\varepsilon }}(c)\rceil \leq \vee \{\lceil path_{\Re
2447: }(c^{\prime })\rceil :c^{\prime }\in T_{\varepsilon }(Q,\Sigma
2448: ),b(c^{\prime })=b(c),\ \ \ \ \ \ \ \ \ \ $$
2449: $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ e(c^{\prime })=e(c)\ {\rm and}\ lb_{\varepsilon }(c^{\prime
2450: })=lb(c)\}.$$ For the case of $|c|=1,$ it is immediate from the
2451: definition of transition relation $\delta ^{\prime }$ in $\Re
2452: ^{-\varepsilon }.$ If $c=c^{\prime }\sigma q,$ then with the
2453: induction hypothesis and Lemmas 2.5 and 2.6, we have
2454: $$\lceil \gamma (atom(\Re ))\rceil \wedge \lceil path_{\Re
2455: ^{-\varepsilon }}(c)\rceil =\lceil \gamma (atom(\Re ))\rceil
2456: \wedge \lceil path_{\Re ^{-\varepsilon }}(c^{\prime })\rceil
2457: \wedge \delta ^{\prime }(e(c^{\prime }),\sigma ,q)$$
2458: $$=\lceil \gamma (atom(\Re ))\rceil \wedge \lceil \gamma (atom(\Re
2459: ))\rceil \wedge \lceil path_{\Re ^{-\varepsilon }}(c^{\prime
2460: })\rceil \wedge \vee _{m,n\geq 0}\delta (e(c^{\prime
2461: }),\varepsilon ^{m}\sigma \varepsilon ^{n},q) $$
2462: $$\leq \vee _{m,n\geq 0}(\lceil \gamma (atom(\Re ))\rceil \wedge
2463: \lceil path_{\Re ^{-\varepsilon }}(c^{\prime })\rceil \wedge
2464: \delta (e(c^{\prime }),\varepsilon ^{m}\sigma \varepsilon
2465: ^{n},q))$$
2466: $$\leq \vee _{m,n\geq 0}[\lceil \gamma (atom(\Re ))\rceil \wedge
2467: \vee \{\lceil path_{\Re }(d^{\prime })\rceil :d^{\prime }\in
2468: T_{\varepsilon }(Q,\Sigma ),b(d^{\prime })=b(c^{\prime
2469: }),e(d^{\prime })=e(c^{\prime })$$ $${\rm and}\ lb_{\varepsilon
2470: }(d^{\prime })=lb(c^{\prime })\}\wedge \delta (e(c^{\prime
2471: }),\varepsilon ^{m}\sigma \varepsilon ^{n},q)]$$
2472: $$\leq \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge \lceil
2473: path_{\Re }(d^{\prime })\rceil \wedge \delta (e(c^{\prime
2474: }),\varepsilon ^{m}\sigma \varepsilon ^{n},q):m,n\geq 0,d^{\prime
2475: }\in T_{\varepsilon }(Q,\Sigma ),$$
2476: $$b(d^{\prime })=b(c^{\prime }),e(d^{\prime })=e(c^{\prime
2477: }),lb_{\varepsilon }(d^{\prime })=lb(c^{\prime })\}.$$
2478: Furthermore, we know that
2479: $$\delta (e(c^{\prime }),\varepsilon ^{m}\sigma \varepsilon
2480: ^{n},q)=\vee \{\delta (e(c^{\prime }),\varepsilon ,p_{1})\wedge
2481: \delta (p_{1},\varepsilon ,p_{2})\wedge ...\wedge \delta
2482: (p_{m-1},\varepsilon ,p_{m})\wedge \delta (p_{m},\sigma ,q_{n})$$
2483: $$\wedge \delta (q_{n},\varepsilon ,q_{n-1})\wedge ...\wedge \delta
2484: (q_{2},\varepsilon ,q_{1})\wedge \delta (q_{1},\varepsilon
2485: ,q):p_{1},...,p_{m},q_{1},...,q_{n}\in Q\}.$$ Again, we use Lemmas
2486: 2.5 and 2.6 and obtain
2487: $$\lceil \gamma (atom(\Re ))\rceil \wedge \lceil path_{\Re
2488: ^{-\varepsilon }}(c)\rceil \leq \vee \{\lceil path_{\Re
2489: }(d^{\prime })\rceil \wedge \delta (e(c^{\prime }),\varepsilon
2490: ,p_{1})\wedge \delta (p_{1},\varepsilon ,p_{2})\wedge ...$$
2491: $$\wedge \delta (p_{m-1},\varepsilon ,p_{m})\wedge \delta
2492: (p_{m},\sigma ,q_{n})\wedge \delta (q_{n},\varepsilon
2493: ,q_{n-1})\wedge ...\wedge \delta (q_{2},\varepsilon ,q_{1})\wedge
2494: \delta (q_{1},\varepsilon ,q):$$ $$m,n\geq 0,d^{\prime }\in
2495: T_{\varepsilon }(Q,\Sigma )\ {\rm with}\ b(d^{\prime
2496: })=b(c^{\prime }),e(d^{\prime })=e(c^{\prime })$$ $${\rm and}\
2497: lb_{\varepsilon }(d^{\prime })=lb(c^{\prime
2498: }),p_{1},...,p_{m},q_{1},...,q_{n}\in Q\}.$$ We put $d=d^{\prime
2499: }\varepsilon p_{1}\varepsilon p_{2}...p_{m-1}\varepsilon
2500: p_{m}\sigma q_{n}\varepsilon q_{n-1}...q_{2}\varepsilon
2501: q_{1}\varepsilon q.$ Then $b(d)=b(d^{\prime })=b(c^{\prime }),$
2502: $e(d)=q=e(c),$ $lb_{\varepsilon }(d)=lb_{\varepsilon }(d^{\prime
2503: })\sigma =lb(c^{\prime })\sigma =lb(c),$ and
2504: $$\lceil path_{\Re }(d)\rceil =\lceil path_{\Re }(d^{\prime
2505: })\rceil \wedge \delta (e(c^{\prime }),\varepsilon ,p_{1})\wedge
2506: \delta (p_{1},\varepsilon ,p_{2})\wedge ...$$
2507: $$\wedge \delta (p_{m-1},\varepsilon ,p_{m})\wedge \delta
2508: (p_{m},\sigma ,q_{n})\wedge \delta (q_{n},\varepsilon
2509: ,q_{n-1})\wedge ...\wedge \delta (q_{2},\varepsilon ,q_{1})\wedge
2510: \delta (q_{1},\varepsilon ,q).$$ Therefore,
2511: $$\lceil \gamma (atom(\Re ))\rceil \wedge \lceil path_{\Re
2512: ^{-\varepsilon }}(c)\rceil \leq \vee \{\lceil path_{\Re }(d)\rceil
2513: :d\in T_{\varepsilon }(Q,\Sigma ),b(d)=b(c),$$ $$e(d)=e(c)\ {\rm
2514: and}\ lb_{\varepsilon }(d)=lb(c)\}$$ and the claim holds for the
2515: case of $|c|=|c^{\prime }|+1.$
2516:
2517: Now it follows from the above claim and Lemmas 2.5 and 2.6 that
2518: $$\lceil \gamma (atom(\Re ))\rceil \wedge \lceil rec_{\Re
2519: ^{-\varepsilon }}(s)\rceil =\lceil \gamma (atom(\Re ))\rceil
2520: \wedge \lceil \gamma (atom(\Re ))\rceil \wedge \vee
2521: \{I(b(c))\wedge T^{\prime }(e(c))$$
2522: $$\wedge \lceil path_{\Re ^{-\varepsilon }}(c)\rceil :c\in
2523: T(Q,\Sigma ),lb(c)=s\}$$
2524: $$\leq \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge I(b(c))\wedge
2525: T^{\prime }(e(c))\wedge \lceil path_{\Re ^{-\varepsilon
2526: }}(c)\rceil :c\in T(Q,\Sigma ),lb(c)=s\}$$
2527: $$\leq \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge I(b(c))\wedge
2528: T^{\prime }(e(c))\wedge \vee \{\lceil path_{\Re }(c^{\prime
2529: })\rceil :c^{\prime }\in T_{\varepsilon }(Q,\Sigma ),b(c^{\prime
2530: })=b(c),$$
2531: $$e(c^{\prime })=e(c)\ {\rm and}\ lb_{\varepsilon }(c^{\prime
2532: })=lb(c)\}:c\in T(Q,\Sigma ),lb(c)=s\}$$
2533: $$\leq \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge I(b(c))\wedge
2534: T^{\prime }(e(c))\wedge \lceil path_{\Re }(c^{\prime })\rceil
2535: :c\in T(Q,\Sigma ),$$
2536: $$c^{\prime }\in T_{\varepsilon }(Q,\Sigma ),b(c^{\prime
2537: })=b(c),e(c^{\prime })=e(c)\ {\rm and}\ lb_{\varepsilon
2538: }(c^{\prime })=lb(c)=s\}$$
2539: $$=\vee \{\lceil \gamma (atom(\Re ))\rceil \wedge I(b(c))\wedge
2540: \vee _{m\geq 0}(T(q)\wedge \delta (e(c),\varepsilon ^{m},q))\wedge
2541: \lceil path_{\Re }(c^{\prime })\rceil :c\in T(Q,\Sigma ),$$
2542: $$c^{\prime }\in T_{\varepsilon }(Q,\Sigma ),b(c^{\prime
2543: })=b(c),e(c^{\prime })=e(c)\ {\rm and}\ lb_{\varepsilon
2544: }(c^{\prime })=lb(c)=s\}$$ $$\leq \vee \{\lceil \gamma (atom(\Re
2545: ))\rceil \wedge I(b(c))\wedge T(q)\wedge \lceil path_{\Re
2546: }(c^{\prime })\rceil \wedge \delta (e(c),\varepsilon ^{m},q):m\geq
2547: 0,c\in T(Q,\Sigma ),$$
2548: $$c^{\prime }\in T_{\varepsilon }(Q,\Sigma ),b(c^{\prime
2549: })=b(c),e(c^{\prime })=e(c)\ {\rm and}\ lb_{\varepsilon
2550: }(c^{\prime })=lb(c)=s\}$$
2551: $$\leq \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge I(b(c))\wedge
2552: T(q)\wedge \lceil path_{\Re }(c^{\prime })\rceil \wedge \delta
2553: (e(c),\varepsilon ,q_{1})\wedge \delta (q_{1},\varepsilon
2554: ,q_{2})\wedge ...$$
2555: $$\wedge \delta (q_{m-2},\varepsilon ,q_{m-1})\wedge \delta
2556: (q_{m-1},\varepsilon q):m\geq 0,c\in T(Q,\Sigma ),c^{\prime }\in
2557: T_{\varepsilon }(Q,\Sigma ),b(c^{\prime })=b(c),$$
2558: $$e(c^{\prime })=e(c),lb_{\varepsilon }(c^{\prime
2559: })=lb(c)=s,q_{1},...,q_{m-1}\in Q\}.$$ If we write $d=c^{\prime
2560: }\varepsilon q_{1}\varepsilon q_{2}...q_{m-2}\varepsilon
2561: q_{m-1}\varepsilon q,$ then $b(d)=b(c^{\prime })=b(c),$ $e(d)=q,$
2562: $lb_{\varepsilon }(d)=lb(c^{\prime })=s$ and
2563: $$\lceil path_{\Re }(d)\rceil =\lceil path_{\Re }(c^{\prime
2564: })\rceil \wedge \delta (e(c),\varepsilon ,q_{1})\wedge \delta
2565: (q_{1},\varepsilon ,q_{2})\wedge ...\wedge \delta
2566: (q_{m-2},\varepsilon ,q_{m-1})\wedge \delta (q_{m-1},\varepsilon
2567: ,q).$$ Thus,
2568: $$\lceil \gamma (atom(\Re ))\rceil \wedge \lceil rec_{\Re
2569: ^{-\varepsilon }}(s)\rceil \leq \vee \{I(b(d))\wedge T(e(d))\wedge
2570: \lceil path_{\Re }(d)\rceil :d\in T_{\varepsilon }(Q,\Sigma
2571: ),lb_{\varepsilon }(d)=s\}$$
2572: $$=\lceil rec_{\Re }(s)\rceil .$$
2573:
2574: For (3), the part from (i) to (ii) is immediate from (2) by noting
2575: that $ \lceil \gamma (atom(\Re ))\rceil =1$ always holds in a
2576: Boolean algebra $\ell .$ Conversely, we demonstrate that (ii)
2577: implies (i). For any $a,b,c\in L,$ consider $\ell -$valued
2578: automaton $\Re
2579: =<\{q_{0},q_{1},...,q_{5}\},\{q_{0}\},\{q_{5}\},\delta >$ with
2580: $\varepsilon - $moves in which $\sigma \in \Sigma ,$ $\delta
2581: (q_{0},\sigma ,q_{1})=a,$ $ \delta (q_{1},\varepsilon
2582: ,q_{2})=\delta (q_{1},\varepsilon ,q_{3})=\delta (q_{4},\sigma
2583: ,q_{5})=1,$ $\delta (q_{2},\varepsilon ,q_{4})=b,$ $\delta
2584: (q_{3},\varepsilon ,q_{4})=c,$ and $\delta $ takes values $0$ for
2585: other arguments (see Figure 4).
2586:
2587: \begin{figure}\centering
2588: \includegraphics{fig_4.eps}\caption{Automaton b} \label{fig 4}
2589: \end{figure}
2590:
2591: By a routine calculation we know that its $\varepsilon -$reduction
2592: is $\Re ^{-\varepsilon
2593: }=<\{q_{0},q_{1},...,q_{5}\},\{q_{0}\},\{q_{5}\},\delta ^{\prime
2594: }>$ where $\delta ^{\prime }(q_{0},\sigma ,q_{1})=\delta ^{\prime
2595: }(q_{0},\sigma ,q_{2})=\delta ^{\prime }(q_{0},\sigma ,q_{3})=a,$
2596: $\delta ^{\prime }(q_{0},\sigma ,q_{4})=(a\wedge b)\vee (a\wedge
2597: c),$ $\delta ^{\prime }(q_{1},\sigma ,q_{5})=b\vee c,$ $\delta
2598: ^{\prime }(q_{2},\sigma ,q_{5})=b,$ $\delta ^{\prime
2599: }(q_{3},\sigma ,q_{5})=c,$ $\delta ^{\prime }(q_{4},\sigma
2600: ,q_{5})=1,$ and $\delta $ takes value $0$ for other arguments (see
2601: Figure 5). Then it follows from (ii) that
2602: $$a\wedge (b\vee c)=[a\wedge (b\vee c)]\vee (a\wedge b)\vee
2603: (a\wedge c)\vee \lbrack (a\wedge b)\vee (a\wedge c)]$$
2604: $$=\lceil rec_{\Re ^{-\varepsilon }}(\sigma \sigma )\rceil $$
2605: $$=\lceil rec_{\Re }(\sigma \sigma )\rceil $$
2606: $$=(a\wedge b)\vee (a\wedge c).$$
2607: This shows that $\ell $ enjoys the distributivity of meet over
2608: union, and it is a Boolean algebra.$\heartsuit$
2609:
2610: \begin{figure}\centering
2611: \includegraphics[width=1\textwidth]{fig_5.eps}\caption{Automaton c} \label{fig 5}
2612: \end{figure}
2613:
2614: \bigskip\
2615:
2616: \textbf{6. Closure Properties of Orthomodular Lattice-Valued
2617: Regularity}
2618:
2619: \smallskip\
2620:
2621: It was shown in the classical automata theory that the class of
2622: regular languages is closed under various operations such as
2623: union, intersection, complement, concatenation, the Kleene
2624: closure, substitution and homomorphism. In this section, we are
2625: going to examine the closure properties of orthomodular
2626: lattice-valued languages under these operations. We first consider
2627: the inverse of an $\ell-$valued language. Let $A\in L^{\Sigma
2628: ^{*}}.$ Then the inverse $A^{-1}\in L^{\Sigma ^{*}}$ of $ A$ is
2629: defined as follows:
2630: $$A^{-1}(\sigma _1...\sigma _m)=A(\sigma _m...\sigma _1)$$ for any
2631: $m\in \omega $ and for any $\sigma _1,...,\sigma _m\in \Sigma .$
2632:
2633: The following proposition shows that both regularity and
2634: commutative regularity are preserved by the inverse operation.
2635:
2636: \smallskip\
2637:
2638: \textbf{Proposition 6.1.} Let $\ell=<L,\leq ,\wedge ,\vee ,\bot
2639: ,0,1>$ be a complete orthomodular lattice, and let
2640: $\longrightarrow $ fulfil the property that $a\leftrightarrow a=1$
2641: for any $a\in L$. Then for any $A\in L^{\Sigma ^{*}},$
2642: $$\stackrel{\ell }{ \models }Reg_\Sigma (A)\leftrightarrow
2643: Reg_\Sigma (A^{-1}),$$ and
2644: $$\stackrel{\ell }{ \models }CReg_\Sigma (A)\leftrightarrow
2645: CReg_\Sigma (A^{-1}).$$
2646:
2647: \smallskip\
2648:
2649: \textbf{Proof}. Noting that $A=(A^{-1})^{-1},$ it suffices to show
2650: that $$ \lceil Reg_\Sigma (A)\rceil \leq \lceil Reg_\Sigma
2651: (A^{-1})\rceil .$$ For any $\ell -$valued automaton $\Re
2652: =(Q,I,T,\delta ),$ we define the inverse of $ \Re $ to be the
2653: $\ell -$valued automaton $\Re ^{-1}=(Q,T,I,\delta ^{-1}),$ where
2654: $\delta ^{-1}(p,\sigma ,q)=\delta (q,\sigma ,p)$ for any $p,q\in
2655: Q$ and $\sigma \in \Sigma .$ Then it is easy to see that $rec_{\Re
2656: ^{-1}}=(rec_\Re )^{-1},$ and furthermore we have
2657: $$\lceil
2658: Reg_\Sigma (A)\rceil =\vee \{\lceil A\equiv rec_\Re \rceil :\Re
2659: \in \mathbf{A}(\Sigma ,\ell )\}$$
2660: $$=\vee \{\lceil A^{-1}\equiv
2661: (rec_\Re )^{-1}\rceil :\Re \in \mathbf{A} (\Sigma ,\ell )\}$$
2662: $$=\vee \{\lceil A^{-1}\equiv rec_{\Re ^{-1}}\rceil :\Re \in
2663: \mathbf{A} (\Sigma ,\ell )\}$$
2664: $$\leq \vee \{\lceil A^{-1}\equiv
2665: rec_\wp \rceil :\wp \in \mathbf{A}(\Sigma ,\ell )\}=\lceil
2666: Rec_\Sigma (A^{-1})\rceil .$$. The proof for commutative
2667: regularity is similar. $\heartsuit$
2668:
2669: \smallskip\
2670:
2671: The commutative regularity is preserved by the complement
2672: operation, but it is not the case for the regularity predicate.
2673:
2674: \smallskip\
2675:
2676: \textbf{Proposition 6.2.} If $\ell =<L,\leq ,\wedge ,\vee ,\perp
2677: ,0,1>$ is a finite orthomodular lattice, and $\rightarrow
2678: =\rightarrow _{3},$ then for any $A\in L^{\Sigma ^{\ast }},$
2679: $$\stackrel{\ell }{\models }CReg_{\Sigma }(A)\rightarrow
2680: CReg_{\Sigma }(A^{c}).$$
2681:
2682: \smallskip\
2683:
2684: \textbf{Proof.} For any $\Re =<Q,I,T,\delta >\in \mathbf{A}(\Sigma
2685: ,\ell ),$ we observe that $\ell ^{\Re
2686: }=<L^{Q},I_{1},\overline{T},\overline{\delta }>$ is an $\ell
2687: -$valued deterministic automaton and only $\overline{T}$ carries
2688: $\ell -$valued information. Then we set $(\ell ^{\Re
2689: })^{c}=<L^{Q},I_{1}, \overline{T}^{c},\overline{\delta }>,$ where
2690: for any $X\in L^{Q},$ $
2691: \overline{T}^{c}(X)=(\overline{T}(X))^{c}.$ It is easy to see that
2692: for all $ s\in \Sigma ^{\ast },$ $rec_{(\ell ^{\Re
2693: })^{c}}(s)=(rec_{(\ell ^{\Re })}(s))^{\perp }.$
2694:
2695: Now by using Theorem 4.1 and Lemmas 2.5 and 2.6 we obtain
2696: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil A\equiv
2697: rec_{\Re }\rceil \leq \lceil \gamma (atom(\Re )\cup r(A)\rceil
2698: \wedge \lceil A\equiv rec_{\Re }\rceil \wedge \lceil rec_{\Re
2699: }\equiv rec_{(\ell ^{\Re })}\rceil $$
2700: $$=\wedge _{s\in \Sigma ^{\ast }}(\lceil \gamma (atom(\Re )\cup
2701: r(A)\rceil \wedge (A(s)\rightarrow rec_{\Re }(s))\wedge (rec_{\Re
2702: }(s)\rightarrow rec_{(\ell ^{\Re })}(s)))\wedge $$
2703: $$\wedge _{s\in \Sigma ^{\ast }}(\lceil \gamma (atom(\Re )\cup
2704: r(A)\rceil \wedge (rec_{(\ell ^{\Re })}(s)\rightarrow rec_{\Re
2705: }(s))\wedge (rec_{\Re }(s)\rightarrow A(s)))$$
2706: $$\leq \wedge _{s\in \Sigma ^{\ast }}(\lceil \gamma (atom(\Re )\cup
2707: r(A)\rceil \wedge (A(s)\rightarrow rec_{(\ell ^{\Re })}(s)))\wedge
2708: $$
2709: $$\wedge _{s\in \Sigma ^{\ast }}(\lceil \gamma (atom(\Re )\cup
2710: r(A)\rceil \wedge (rec_{(\ell ^{\Re })}(s)\rightarrow A(s))).$$
2711: Then Lemma 2.11(2) yields
2712: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil A\equiv
2713: rec_{\Re }\rceil \leq \wedge _{s\in \Sigma ^{\ast }}(rec_{(\ell
2714: ^{\Re })}^{\perp }(s)\rightarrow A^{\perp }(s))\wedge \wedge
2715: _{s\in \Sigma ^{\ast }}(A^{\perp }(s)\rightarrow rec_{(\ell ^{\Re
2716: })}^{\perp }(s))$$
2717: $$=\wedge _{s\in \Sigma ^{\ast }}(A^{\perp }(s)\leftrightarrow
2718: rec_{(\ell ^{\Re })}^{\perp }(s))$$
2719: $$=\wedge _{s\in \Sigma ^{\ast }}(A^{\perp }(s)\leftrightarrow
2720: rec_{(\ell ^{\Re })^{c}}(s))$$
2721: $$=\lceil A^{c}\equiv rec_{(\ell ^{\Re })^{c}}\rceil .$$
2722:
2723: In addition, we have $$\lceil \gamma (atom(\Re )\cup r(A)\rceil
2724: \leq \lceil \gamma (atom(\Re )\cup r(A^{c})\rceil $$ from Lemma
2725: 2.5. Therefore,
2726: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil A\equiv
2727: rec_{\Re }\rceil \leq \lceil \gamma (atom(\Re )\cup r(A^{c})\rceil
2728: \wedge \lceil A^{c}\equiv rec_{(\ell ^{\Re })^{c}}\rceil $$
2729: $$\leq \lceil CReg_{\Sigma }(A^{c})\rceil .$$
2730:
2731: Finally, since $\Re $ is allowed to be arbitrary, it follows that
2732: $$\lceil CReg_{\Sigma }(A)\rceil =\vee _{\Re \in
2733: \mathbf{A}(\Sigma ,\ell )}\lceil \gamma (atom(\Re )\cup r(A)\rceil
2734: \wedge \lceil A\equiv rec_{\Re }\rceil $$
2735: $$\leq \lceil CReg_{\Sigma }(A^{c})\rceil .\heartsuit$$
2736:
2737: \smallskip\
2738:
2739: We now turn to deal with the union of two $\ell-$valued language.
2740: Let $\Re =<Q_{A},I_{A},T_{A},\delta _{A})$ and $\wp
2741: =<Q_{B},I_{B},T_{B},\delta _{B}>\in \mathbf{A}(\Sigma ,\ell )$ be
2742: two $\ell - $valued automata over $\Sigma .$ We assume that
2743: $Q_{A}\cap Q_{B}=\phi .$ Then the (disjoint) union $\Re \cup \wp $
2744: of $\Re $ and $\wp $ is defined to be $\Im
2745: =(Q_{C},I_{C},T_{C},\delta _{C}),$ where:
2746:
2747: (i) $Q_{C}=Q_{A}\cup Q_{B};$
2748:
2749: (ii) $I_{C}=I_{A}\cup I_{B};$
2750:
2751: (iii) $T_{C}=T_{A}\cup T_{B};$ and
2752:
2753: (iv) $\delta _{C}:Q_{C}\times \Sigma \times Q_{C}\longrightarrow
2754: L$ is given as follows: for any $p,q\in Q_{C}$ and $\sigma \in
2755: \Sigma ,$
2756: $$\delta _{C}(p,\sigma ,q)=\left\{
2757: \begin{array}{c}
2758: \delta _{A}(p,\sigma ,q)\ {\rm if }\ p,q\in Q_{A}, \\
2759: \delta _{B}(p,\sigma ,q)\ {\rm if }\ p,q\in Q_{B}, \\
2760: 0\ {\rm otherwise.}
2761: \end{array}
2762: \right. $$
2763:
2764: The following proposition describes the recognizability of the
2765: union of two $\ell-$valued automata. As in the classical theory, a
2766: word $s$ in $\Sigma^{\ast}$ is recognized by the union of two
2767: $\ell-$valued automata if and only if $s$ is recognized by one of
2768: them.
2769:
2770: \smallskip\
2771:
2772: \textbf{Proposition 6.3.} Let $\ell=<L,\leq ,\wedge ,\vee ,\bot
2773: ,0,1>$ be a complete orthomodular lattice. If the implication
2774: operator $\longrightarrow $ satisfies that $a\leftrightarrow a=1$
2775: for any $a\in L,$ then for any $\Re ,\wp \in \mathbf{A(\Sigma
2776: ,\ell )}$ and for any $s\in \Sigma ^{\ast },$
2777: $$\stackrel{\ell }{ \models } rec_{\Re \cup \wp
2778: }(s)\longleftrightarrow rec_{\Re }(s)\vee rec_{\wp }(s).$$
2779:
2780: \smallskip\
2781:
2782: \textbf{Proof}. Let $s=\sigma _{1}...\sigma _{k}.$ Then
2783: $$\lceil rec_{\Re \cup \wp }(s)\rceil =\vee \{(I_{A}\cup
2784: I_{B})(q_{0})\wedge (T_{A}\cup T_{B})(q_{k})\wedge \wedge
2785: _{i=0}^{k-1}\delta _{A\cup B}(q_{i},\sigma
2786: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in Q_{A}\cup Q_{B}\}$$
2787: $$=[\vee \{(I_{A}\cup I_{B})(q_{0})\wedge (T_{A}\cup
2788: T_{B})(q_{k})\wedge \wedge _{i=0}^{k-1}\delta _{A\cup
2789: B}(q_{i},\sigma _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in
2790: Q_{A}\}]$$
2791: $$\ \ \ \ \vee \lbrack \vee \{(I_{A}\cup I_{B})(q_{0})\wedge (T_{A}\cup
2792: T_{B})(q_{k})\wedge \wedge _{i=0}^{k-1}\delta _{A\cup
2793: B}(q_{i},\sigma _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in
2794: Q_{B}\}]$$
2795: $$\ \ \ \ \ \ \vee \lbrack \vee \{(I_{A}\cup I_{B})(q_{0})\wedge (T_{A}\cup
2796: T_{B})(q_{k})\wedge \wedge _{i=0}^{k-1}\delta _{A\cup
2797: B}(q_{i},\sigma _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in Q_{A}\cup
2798: Q_{B},$$ $${\rm and\ there\ are}\ i,j\ {\rm such\ that}\ 0\leq
2799: i,j\leq k\ {\rm and}\ q_{i}\in Q_{A}\ {\rm and}\ q_{j}\in
2800: Q_{B}\}].$$ From the definition of $\Re \cup \wp ,$ we know that
2801: for any $ q_{0},q_{1},...,q_{k}\in Q_{A},$ $$(I_{A}\cup
2802: I_{B})(q_{0})=I_{A}(q_{0}),$$ $$(T_{A}\cup
2803: T_{B})(q_{k})=T_{A}(q_{k}),$$
2804: $$\wedge _{i=0}^{k-1}\delta _{A\cup B}(q_{i},\sigma
2805: _{i+1},q_{i+1})=\wedge _{i=0}^{k-1}\delta _{A}(q_{i},\sigma
2806: _{i+1},q_{i+1}),$$ and for any $q_{0},q_{1},...,q_{k}\in Q_{B},$
2807: $$(I_{A}\cup I_{B})(q_{0})=I_{B}(q_{0}),$$
2808: $$(T_{A}\cup T_{B})(q_{k})=T_{B}(q_{k}),$$
2809: $$\wedge _{i=0}^{k-1}\delta _{A\cup B}(q_{i},\sigma
2810: _{i+1},q_{i+1})=\wedge _{i=0}^{k-1}\delta _{B}(q_{i},\sigma
2811: _{i+1},q_{i+1}).$$ If $q_{0},q_{1},...,q_{k}\in Q_{A}\cup Q_{B},$
2812: and there are $i,j$ such that $0\leq i,j\leq k$ and $q_{i}\in
2813: Q_{A}$ and $q_{j}\in Q_{B},$ then we can find some $m\in
2814: \{0,1,...,k-1\}$ such that $q_{m}\in Q_{A}$ and $q_{m+1}\in
2815: Q_{B},$ or $q_{m}\in Q_{B}$ and $q_{m+1}\in Q_{A}.$ Then $\delta
2816: _{A\cup B}(q_{m},\sigma _{m+1},q_{m+1})=0,$ and $$\wedge
2817: _{i=0}^{k-1}\delta _{A\cup B}(q_{i},\sigma _{i+1},q_{i+1})=0.$$
2818: Therefore, it follows that
2819: $$\lceil rec_{\Re \cup \wp }(s)\rceil =[\vee \{I_{A}(q_{0})\wedge
2820: T_{A}(q_{k})\wedge \wedge _{i=0}^{k-1}\delta _{A}(q_{i},\sigma
2821: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in Q_{A}\}]$$
2822: $$\vee \lbrack \vee \{I_{B}(q_{0})\wedge T_{B}(q_{k})\wedge \wedge
2823: _{i=0}^{k-1}\delta _{B}(q_{i},\sigma
2824: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in Q_{B}\}]$$
2825: $$=\lceil rec_{\Re }(s)\vee rec_{\wp }(s)\rceil .\heartsuit $$
2826:
2827: \smallskip\
2828:
2829: The following corollary slightly generalizes Example 3.1.
2830:
2831: \smallskip\
2832:
2833: \textbf{Corollary 6.4}. If $Range(A)=\{A(s):s\in \Sigma^{\ast}\}$
2834: is finite, and $A_\lambda =\{s\in\Sigma^{\ast}:A(s)\geq \lambda\}$
2835: is a regular language (in classical automata theory) for every
2836: $\lambda\in Range(A)$, then
2837: $$\stackrel{\ell }{ \models }Reg_\Sigma (A).$$
2838:
2839: \smallskip\
2840:
2841: \textbf{Proof.} Suppose that
2842: $Range(A)=\{\lambda_1,...,\lambda_n\}.$ Then it is easy to see
2843: that
2844: $$A=\cup_{i=1}^{n} \lambda_i A_{\lambda_i}.$$ From Example 3.3 we
2845: know that there exists an $\ell-$valued automaton $\Re_i$ such
2846: that $rec_{\Re_i}=\lambda_i A_{\lambda_i}$ for each $i\leq n$.
2847: Thus, by proposition 6.3 we obtain $$rec_{\cup_{i=1}^{n} \Re_i}
2848: =\cup_{i=1}^{n} \lambda_i A_{\lambda_i}=A$$ and complete the
2849: proof. $\heartsuit$
2850:
2851: \smallskip\
2852:
2853: We now consider the product of two $\ell-$valued automata. Let
2854: $\Re =<Q_{A},I_{A},T_{A},\delta _{A})$ and $\wp
2855: =<Q_{B},I_{B},T_{B},\delta _{B}>\in \mathbf{A}(\Sigma ,\ell )$ be
2856: two $\ell -$valued automata over $ \Sigma .$ Then their product
2857: $\Re \times \wp $ is defined to be $\Im =(Q_{C},I_{C},T_{C},\delta
2858: _{C}),$ where:
2859:
2860: \smallskip\
2861:
2862: (i) $Q_{C}=Q_{A}\times Q_{B};$
2863:
2864: (ii) $I_{C}=I_{A}\times I_{B};$
2865:
2866: (iii) $T_{C}=T_{A}\times T_{B};$ and
2867:
2868: (iv) $\delta _{C}:Q_{C}\times \Sigma \times Q_{C}\longrightarrow
2869: L$ and for any $p_{a},q_{a}\in Q_{A},$ $p_{b},q_{b}\in Q_{B}$ and
2870: $\sigma \in \Sigma ,$
2871: $$\delta _{C}((p_{a},p_{b}),\sigma ,(q_{a},q_{b}))=\delta
2872: _{A}(p_{a},\sigma ,q_{a})\wedge \delta _{B}(p_{b},\sigma
2873: ,q_{b}).$$
2874:
2875: \smallskip\
2876:
2877: It is well-known in the classical automata theory that the
2878: language accepted by the union of two automata is the union of the
2879: languages accepted by these two automata, and the language
2880: accepted by the product of two automata is the intersection of the
2881: languages accepted by the factor automata. Proposition 6.3 shows
2882: that the conclusion about the union of two automata can be
2883: generalized into the theory of automata based on quantum logic
2884: without appealing any additional condition. One may naturally
2885: expect that the conclusion for product of automata can also be
2886: easily generalized into the framework of quantum logic. However,
2887: the case for the product of two automata is much more complicated,
2888: and the following proposition tells us that in order to make the
2889: above conclusion about product of automata still valid in the
2890: automata theory based on quantum logic, a certain commutativity is
2891: necessary to be added on the basic actions of the factor automata.
2892:
2893: \smallskip\
2894:
2895: \textbf{Proposition 6.5.} Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
2896: ,0,1>$ be a complete orthomodular lattice.
2897:
2898: (1) For any $\Re ,\wp \in \mathbf{A}(\Sigma ,\ell ),$ and for any
2899: $s\in \Sigma ^{\ast },$
2900: $$\stackrel{\ell }{\models }rec_{\Re \times \wp }(s)\longrightarrow
2901: rec_{\Re }(s)\wedge rec_{\wp }(s).$$
2902:
2903: (2) For any $\Re ,\wp \in \mathbf{A}(\Sigma ,\ell ),$ and for any
2904: $s\in \Sigma ^{\ast },$
2905: $$\stackrel{\ell }{\models }\gamma (atom(\Re )\cup atom(\wp ))\wedge
2906: rec_{\Re }(s)\wedge rec_{\wp }(s)\longrightarrow rec_{\Re \times
2907: \wp }(s),$$ and in particular if $\longrightarrow =\longrightarrow
2908: _{3},$ then
2909: $$\stackrel{\ell }{\models}\gamma (atom(\Re )\cup atom(\wp
2910: ))\longrightarrow (rec_{\Re }(s)\wedge rec_{\wp
2911: }(s)\longleftrightarrow rec_{\Re \times \wp }(s)).$$
2912:
2913: (3) The following two statements are equivalent:
2914:
2915: \ \ \ \ \ \ (i) $\ell $ is a Boolean algebra.
2916:
2917: \ \ \ \ \ \ (ii) for all $\Re ,\wp \in \mathbf{A}(\Sigma ,\ell ),$
2918: and for all $s\in \Sigma ^{\ast },$
2919: $$\stackrel{\ell }{\models }rec_{\Re }(s)\wedge rec_{\wp
2920: }(s)\longleftrightarrow rec_{\Re \times \wp }(s).$$
2921:
2922: \smallskip\
2923:
2924: \textbf{Proof}. We have directly
2925: $$\lceil rec_{\Re \times \wp }(s)\rceil =\vee \{(I_{A}\times
2926: I_{B})(q_{a0},q_{b0})\wedge (T_{A}\times
2927: T_{B})(q_{ak},q_{bk})\wedge \wedge _{i=0}^{k-1}\delta _{A\times
2928: B}((q_{ai},q_{bi}),$$ $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
2929: \ \sigma _{i+1},(q_{a(i+1)},q_{b(i+1)})):
2930: q_{a0},q_{a1},...,q_{ak}\in Q_{A}\ {\rm and}\
2931: q_{b0},q_{b1},...,q_{bk}\in Q_{B}\}$$
2932: $$=\vee \{I_{A}(q_{a0})\wedge I_{B}(q_{b0})\wedge
2933: T_{A}(q_{ak})\wedge T_{B}(q_{bk})\wedge \wedge _{i=0}^{k-1}\delta
2934: _{A}(q_{ai},\sigma _{i+1},q_{a(i+1)})\wedge$$
2935: $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \wedge
2936: _{i=0}^{k-1}\delta _{B}(q_{bi},\sigma _{i+1},q_{b(i+1)}):
2937: q_{a0},q_{a1},...,q_{ak}\in Q_{A}\ {\rm and}\
2938: q_{b0},q_{b1},...,q_{bk}\in Q_{B}\}, $$ and
2939: $$\lceil rec_{\Re }(s)\wedge rec_{\wp }(s)\rceil =[\vee
2940: \{I_{A}(q_{0})\wedge T_{A}(q_{k})\wedge \wedge _{i=0}^{k-1}\delta
2941: _{A}(q_{i},\sigma _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in
2942: Q_{A}\}]$$
2943: $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \wedge \lbrack \vee \{I_{B}(q_{0})\wedge
2944: T_{B}(q_{k})\wedge \wedge _{i=0}^{k-1}\delta _{B}(q_{i},\sigma
2945: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{k}\in Q_{B}\}]$$ from the
2946: definitions of product and recognizability of $\ell -$valued
2947: automata. It is clear that $$\lceil rec_{\Re \times \wp }(s)\rceil
2948: \leq \lceil rec_{\Re }(s)\wedge rec_{\wp }(s)\rceil .$$ This
2949: indicates that (1) is true. By using Lemmas 2.4(2), 2.5 and 2.6
2950: twice, we can deduce that $$\lceil \gamma (atom(\Re )\cup atom(\wp
2951: ))\wedge rec_{\Re }(s)\wedge rec_{\wp }(s)\rceil \leq \lceil
2952: rec_{\Re \times \wp }(s)\rceil .$$ Thus, (2) is proved. The first
2953: part of (3) that (i) implies (ii) is immediately derivable from
2954: (2) because we have $\lceil \gamma (atom(\Re )\cup atom(\wp
2955: ))\rceil =1$ provided $\ell $ is a Boolean algebra. Conversely, we
2956: show that (ii) implies (i) by constructing two $\ell -$valued
2957: automata and examining the behavior of their product. For any
2958: $a,b,c\in L,$ we choose some $\sigma _{0}\in \Sigma $ and set
2959: $$\Re =(\{p\},\{p\},\{p\},\delta _{A}),$$ where $\delta
2960: _{A}(p,\sigma ,p)=a$ if $\sigma =\sigma _{0}$ and $0$ otherwise,
2961: and $$\wp =(\{q,r,s\},\{q\},\{r,s\},\delta _{B}),$$ where $\delta
2962: _{B}(x,\sigma ,y)=b$ if $x=q,$ $y=r,$ and $\sigma =\sigma _{0};$
2963: $c$ if $x=q,$ $y=s,$ and $\sigma =\sigma _{0},$ $0$ otherwise.
2964: Then $\Re ,\wp \in \mathbf{A}(\Sigma ,\ell ),$ and it is easy to
2965: see that
2966: $$\Re \times \wp
2967: =(\{(p,q),(p,r),(p,s)\},\{(p,q)\},\{(p,r),(p,s)\},\delta _{A\times
2968: B}),$$ where $\delta _{A\times B}((p,x),\sigma ,(p,y))=a\wedge b$
2969: if $x=q,$ $y=r$ and $\sigma =\sigma _{0};$ $a\wedge c$ if $x=q,$
2970: $y=s$ and $\sigma =\sigma _{0};$ and $0$ otherwise (see Figure 6).
2971: Furthermore, by a routine calculation we have
2972: $$\lceil rec_{\Re }(\sigma _{0})\rceil =a,$$
2973: $$\lceil rec_{\wp }(\sigma _{0})\rceil =b\vee c,\ {\rm and}$$
2974: $$\lceil rec_{\Re \times \wp }(\sigma _{0})\rceil =(a\wedge b)\vee
2975: (a\wedge c).$$ Therefore, with (ii) we finally obtain
2976: $$a\wedge b\vee c=\lceil rec_{\Re }(\sigma _{0})\rceil \wedge
2977: \lceil rec_{\wp }(\sigma _{0})\rceil $$ $$=\lceil rec_{\Re \times
2978: \wp }(\sigma _{0})\rceil =(a\wedge b)\vee (a\wedge c).\heartsuit$$
2979:
2980: \begin{figure}\centering
2981: \includegraphics[width=0.5\textwidth]{fig_6.eps}\caption{Automaton d} \label{fig 6}
2982: \end{figure}
2983:
2984: \smallskip\
2985:
2986: To prove the closure property of orthomodular lattice-valued
2987: regularity under the concatenation operation of languages, we
2988: propose the concept of concatenation of two orthomodular
2989: lattice-valued automata. Suppose that $\Re
2990: _{1}=<Q_{1},I_{1},T_{1},\delta _{1}>,$ $\Re
2991: _{2}=<Q_{2},I_{2},T_{2},\delta _{2}>\in \mathbf{A}(\Sigma ,\ell )$
2992: be two $ \ell -$valued automata, and $Q_{1}\cap Q_{2}=\phi .$ We
2993: define the concatenation of $\Re _{1}$ and $\Re _{2}$ to be $\ell
2994: -$valued automaton $ \Re _{1}\Re _{2}=<Q_{1}\cup
2995: Q_{2},I_{1},T_{2},\delta _{2}>$ with $ \varepsilon -$moves, where
2996: $\delta :Q\times (\Sigma \cup \{\varepsilon \})\times Q\rightarrow
2997: L$ is given by
2998: $$\delta (p,\sigma ,q)=\left\{
2999: \begin{array}{c}
3000: \delta _{1}(p,\sigma ,q)\ {\rm if }\ p,q\in Q_{1}\ {\rm and }\
3001: \sigma \neq
3002: \varepsilon \\
3003: \delta _{2}(p,\sigma ,q)\ {\rm if }\ p,q\in Q_{2}\ {\rm and }\
3004: \sigma \neq
3005: \varepsilon \\
3006: T_{1}(p)\wedge I_{2}(q)\ {\rm if }\ p\in Q_{1},q\in Q_{2}\ {\rm
3007: and }\ \sigma
3008: =\varepsilon \\
3009: 0\ {\rm otherwise.}
3010: \end{array}
3011: \right. $$
3012:
3013: The following proposition clarifies the relation between the
3014: language recognized by the concatenation of two orthomodular
3015: lattice-valued automata and the concatenation of the languages
3016: recognized by the two automata.
3017:
3018: \smallskip\
3019:
3020: \textbf{Proposition 6.6}. Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
3021: ,0,1>$ be an orthomodular lattice and $\rightarrow $ fulfil the
3022: Birkhoff-von Neumann requirement.
3023:
3024: (1) For all $\Re _{1},\Re _{2}\in \mathbf{A}(\Sigma ,\ell ),$ and
3025: for each $ s\in \Sigma ^{\ast },$
3026: $$\stackrel{\ell }{\models }rec_{\Re _{1}\Re _{2}}(s)\rightarrow
3027: (\exists s_{1},s_{2}\in \Sigma ^{\ast })(s_1 s_2=s\wedge rec_{\Re
3028: _{1}}(s_{1})\wedge rec_{\Re _{2}}(s_{2})).$$
3029:
3030: (2) For all $\Re _{1},\Re _{2}\in \mathbf{A}(\Sigma ,\ell ),$ and
3031: for each $ s\in \Sigma ^{\ast },$
3032: $$\stackrel{\ell }{\models }\gamma (atom(\Re _{1})\cup atom(\Re
3033: _{2}))\wedge (\exists s_{1},s_{2}\in \Sigma ^{\ast })(s_1
3034: s_2=s\wedge $$ $$rec_{\Re _{1}}(s_{1})\wedge rec_{\Re
3035: _{2}}(s_{2}))\rightarrow rec_{\Re _{1}\Re _{2}}(s),$$ and if
3036: $\rightarrow =\rightarrow _{3}$ then
3037: $$\stackrel{\ell }{\models }\gamma (atom(\Re _{1})\cup atom(\Re
3038: _{2}))\rightarrow (rec_{\Re _{1}\Re _{2}}(s)\leftrightarrow
3039: (\exists s_{1},s_{2}\in \Sigma ^{\ast })$$ $$(s_1 s_2=s\wedge
3040: rec_{\Re _{1}}(s_{1})\wedge rec_{\Re _{2}}(s_{2}))).$$
3041:
3042: (3) The following two statements are equivalent:
3043:
3044: \ \ \ \ \ \ (i) $\ell $ is a Boolean algebra;
3045:
3046: \ \ \ \ \ \ (ii) for all $\Re _{1},\Re _{2}\in \mathbf{A}(\Sigma
3047: ,\ell ),$ and for each $ s\in \Sigma ^{\ast },$
3048: $$\stackrel{\ell }{\models }rec_{\Re _{1}\Re _{2}}(s)\leftrightarrow
3049: (\exists s_{1},s_{2}\in \Sigma ^{\ast })(s_1 s_2=s\wedge rec_{\Re
3050: _{1}}(s_{1})\wedge rec_{\Re _{2}}(s_{2})).$$
3051:
3052: \smallskip\
3053:
3054: \textbf{Proof}. (1) For any $q_{0},q_{1},...,q_{m}\in Q_{1}\cup
3055: Q_{2},$ $ \sigma _{1},...,\sigma _{m}\in \Sigma \cup \{\varepsilon
3056: \}$ with $\sigma _{1}...\sigma _{m}=s$ (note that it is possible
3057: that $|s|<m$ since $\sigma _{1},...,\sigma _{m}$ may contain
3058: $\varepsilon$'s), if
3059: $$I_{1}(q_{0})\wedge T_{2}(q_{m})\wedge \wedge _{i=1}^{m}\delta
3060: (q_{i-1},\sigma _{i},q_{i})>0,$$ then there exists $j\leq m$ such
3061: that $\sigma _{j}=\varepsilon ,$ $\sigma _{i}\neq \varepsilon $
3062: $(i\neq j),$ $q_{0},...,q_{j-1}\in Q_{1},$ $ q_{j},...,q_{m}\in
3063: Q_{2}.$ Thus, $s=\sigma _{1}...\sigma _{j-1}\sigma _{j+1}...\sigma
3064: _{n},$ and
3065: $$I_{1}(q_{0})\wedge T_{2}(q_{m})\wedge \wedge _{i=1}^{m}\delta
3066: (q_{i-1},\sigma _{i},q_{i})=I_{1}(q_{0})\wedge T_{2}(q_{m})\wedge
3067: \wedge _{i=1}^{j-1}\delta _{1}(q_{i-1},\sigma _{i},q_{i})$$
3068: $$\wedge T_{1}(q_{j-1})\wedge I_{2}(q_{j})\wedge \wedge
3069: _{i=j+1}^{m}\delta _{2}(q_{i-1},\sigma _{i},q_{i})$$
3070: $$=[I_{1}(q_{0})\wedge T_{1}(q_{j-1})\wedge \wedge
3071: _{i=1}^{j-1}\delta _{1}(q_{i-1},\sigma _{i},q_{i})]\wedge \lbrack
3072: I_{2}(q_{j})\wedge T_{2}(q_{m})\wedge \wedge _{i=j+1}^{m}\delta
3073: _{2}(q_{i-1},\sigma _{i},q_{i})] $$
3074: $$\leq rec_{\Re _{1}}(\sigma _{1}...\sigma _{j-1})\wedge rec_{\Re
3075: _{2}}(\sigma _{j+1}...\sigma _{m})$$
3076: $$\leq \vee \{rec_{\Re _{1}}(s_{1})\wedge rec_{\Re
3077: _{2}}(s_{2}):s_{1}s_{2}=s\}.$$
3078:
3079: (2) First, we can use Lemmas 2.5 and 2.6 to derive that
3080: $$\lceil \gamma (atom(\Re _{1})\cup atom(\Re _{2}))\wedge (\exists
3081: s_{1},s_{2}\in \Sigma ^{\ast })(s_{1}s_{2}=s\wedge rec_{\Re
3082: _{1}}(s_{1})\wedge rec_{\Re _{2}}(s_{2}))\rceil $$
3083: $$=\lceil \gamma (atom(\Re _{1})\cup atom(\Re _{2}))\rceil \wedge
3084: \vee _{s_{1}s_{2}=s}(rec_{\Re _{1}}(s_{1})\wedge rec_{\Re
3085: _{2}}(s_{2}))$$
3086: $$\leq \vee _{s_{1}s_{2}=s}(\lceil \gamma (atom(\Re _{1})\cup
3087: atom(\Re _{2}))\rceil \wedge rec_{\Re _{1}}(s_{1})\wedge rec_{\Re
3088: _{2}}(s_{2})).$$
3089:
3090: For any $s_{1},s_{2}\in \Sigma ^{\ast }$ with $s_{1}s_{2}=s,$ we
3091: use Lemmas 2.5 and 2.6 again, and this yields
3092: $$\lceil \gamma (atom(\Re _{1})\cup atom(\Re _{2}))\rceil \wedge
3093: rec_{\Re _{1}}(s_{1})\wedge rec_{\Re _{2}}(s_{2})=\lceil \gamma
3094: (atom(\Re _{1})\cup atom(\Re _{2}))\rceil \wedge $$
3095: $$\vee _{lb(c_{1})=s_{1}}(I_{1}(b(c_{1}))\wedge
3096: T_{1}(e(c_{1}))\wedge \lceil path_{\Re _{1}}(s_{1})\rceil )\wedge
3097: \vee _{lb(c_{2})=s_{2}}(I_{2}(b(c_{2}))\wedge
3098: T_{2}(e(c_{2}))\wedge \lceil path_{\Re _{2}}(s_{2})\rceil )$$
3099: $$\leq \vee
3100: _{lb(c_{1})=s_{1},lb(c_{2})=s_{2}}(I_{1}(b(c_{1}))\wedge
3101: T_{1}(e(c_{1}))\wedge \lceil path_{\Re _{1}}(s_{1})\rceil \wedge
3102: I_{2}(b(c_{2}))\wedge T_{2}(e(c_{2}))\wedge \lceil path_{\Re
3103: _{2}}(s_{2})\rceil ).$$
3104:
3105: Furthermore, for any $c_{1}=p_{0}\sigma _{1}p_{1}...p_{m-1}\sigma
3106: _{m}p_{m}$ and $c_{2}=q_{0}\tau _{1}q_{1}...q_{n-1}\tau _{n}q_{n}$
3107: with $s_{1}=\sigma _{1}...\sigma _{m}$ and $s_{2}=\tau _{1}...\tau
3108: _{n},$
3109: $$I_{1}(b(c_{1}))\wedge T_{1}(e(c_{1}))\wedge \lceil path_{\Re
3110: _{1}}(s_{1})\rceil \wedge I_{2}(b(c_{2}))\wedge
3111: T_{2}(e(c_{2}))\wedge \lceil path_{\Re _{2}}(s_{2})\rceil =$$
3112: $$I_{1}(p_{0})\wedge T_{2}(q_{m})\wedge \wedge _{i=1}^{m}\delta
3113: _{1}(p_{i-1},\sigma _{i},p_{i})\wedge T_{1}(p_{m})\wedge
3114: I_{2}(q_{0})\wedge \wedge _{j=1}^{n}\delta _{2}(q_{j-1},\tau
3115: _{j},q_{j})$$
3116: $$=I_{1}(p_{0})\wedge T_{2}(q_{m})\wedge \lceil path_{\Re _{1}\Re
3117: _{2}}(p_{0}\sigma _{1}p_{1}...p_{m-1}\sigma _{m}p_{m}\varepsilon
3118: q_{0}\tau _{1}q_{1}...q_{n-1}\tau _{n}q_{n})\rceil $$
3119: $$\leq rec_{\Re _{1}\Re _{2}}(s).$$
3120:
3121: (3) The part that (i) implies (ii) is a simple corollary of (2).
3122: Conversely, it suffices to show that $\ell $ enjoys
3123: distributivity; that is, for any $ a,b,c\in L,$ $a\wedge (b\vee
3124: c)=(a\wedge b)\vee (a\wedge c).$ Let $\Re
3125: _{1}=<\{p_{0},p_{1}\},\{p_{0}\},\{p_{1}\},\delta _{1}>$ and $\Re
3126: _{2}=<\{q_{0},q_{1},q_{2}\},\{q_{0}\},\{q_{1},q_{2}\},\delta
3127: _{2}>,$ where $ \delta _{1}(p_{0},\sigma ,p_{1})=a,$ $\delta
3128: _{2}(q_{0},\sigma ,q_{1})=b,$ $ \delta _{2}(q_{0},\sigma
3129: ,q_{2})=c,$ and $\delta _{1},$ $\delta _{2}$ take value $0$ for
3130: other arguments (see Figure 7). Then it follows that
3131: $$a\wedge (b\vee c)=\lceil (\exists s_{1},s_{2}\in \Sigma ^{\ast
3132: })(s_{1}s_{2}=\sigma \sigma \wedge rec_{\Re _{1}}(s_{1})\wedge
3133: rec_{\Re _{2}}(s_{2}))\rceil $$
3134: $$=\lceil rec_{\Re _{1}\Re _{2}}(\sigma \sigma )\rceil $$
3135: $$=(a\wedge b)\vee (a\wedge c).\heartsuit$$
3136:
3137: \begin{figure}\centering
3138: \includegraphics[width=1\textwidth]{fig_7.eps}\caption{Automaton e} \label{fig 7}
3139: \end{figure}
3140:
3141: \smallskip\
3142:
3143: We now turn to consider the Kleene closure of an orthomodular
3144: lattice-valued language. For this purpose, we need to introduce
3145: the fold construction of an orthomodular lattice-valued automaton.
3146: Let $\Re =<Q,I,T,\delta >\in \mathbf{A}(\Sigma ,\ell )$ be an
3147: $\ell -$valued automaton, and let $q_{0}\notin Q$ be a new state.
3148: We define the fold of $ \Re $ to be $\ell -$valued automaton $\Re
3149: ^{\ast }=<Q\cup \{q_{0}\},\{q_{0}\},T\cup \{q_{0}\},\delta ^{\ast
3150: }>$ with $\varepsilon -$ moves, where
3151: $$\delta ^{\ast }:(Q\cup \{q_{0}\})\times (\Sigma \cup
3152: \{\varepsilon \})\times (Q\cup \{q_{0}\})\rightarrow L$$ is given
3153: by
3154: $$\delta ^{\ast }(p,\sigma ,q)=\left\{
3155: \begin{array}{c}
3156: I(q)\ {\rm if }\ p=q_{0}\ {\rm and }\ \sigma =\varepsilon , \\
3157: \delta (p,\sigma ,q)\ {\rm if }\ p,q\in Q\ {\rm and }\ \sigma \neq
3158: \varepsilon ,
3159: \\
3160: T(p)\wedge I(q)\ {\rm if }\ p,q\in Q\ {\rm and }\ \sigma =\varepsilon , \\
3161: 0\ {\rm otherwise.}
3162: \end{array}
3163: \right. $$
3164:
3165: The language accepted by the fold of an orthomodular
3166: lattice-valued automaton is then clearly presented by the
3167: following proposition.
3168:
3169: \smallskip\
3170:
3171: \textbf{Proposition 6.7}. Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
3172: ,0,1>$ be an orthomodular lattice, and let $\rightarrow $ enjoy
3173: the Birkhoff-von Neumann requirement.
3174:
3175: (1) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and for all $s\in
3176: \Sigma ^{\ast },$
3177: $$\stackrel{\ell }{\models }rec_{\Re ^{\ast }}(s)\rightarrow (\exists
3178: m\geq 0,s_{1},...,s_{m}\in \Sigma ^{\ast })(s_{1}...s_{m}=s\wedge
3179: \wedge _{i=1}^{m}rec_{\Re }(s_{i})).$$
3180:
3181: (2) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and for each $s\in
3182: \Sigma ^{\ast },$
3183: $$\stackrel{\ell }{\models }\gamma (atom(\Re ))\wedge (\exists m\geq
3184: 0,s_{1},...,s_{m}\in \Sigma ^{\ast })(s_{1}...s_{m}=s\wedge \wedge
3185: _{i=1}^{m}rec_{\Re }(s_{i}))\rightarrow rec_{\Re ^{\ast }}(s),$$
3186: and in particular if $\rightarrow =\rightarrow _{3},$ then
3187: $$\stackrel{\ell }{\models }\gamma (atom(\Re ))\rightarrow (rec_{\Re
3188: ^{\ast }}(s)\leftrightarrow (\exists m\geq 0,s_{1},...,s_{m}\in
3189: \Sigma ^{\ast })(s_{1}...s_{m}=s\wedge \wedge _{i=1}^{m}rec_{\Re
3190: }(s_{i}))).$$
3191:
3192: (3) The following two statements are equivalent:
3193:
3194: \ \ \ \ \ \ (i) $\ell $ is a Boolean algebra;
3195:
3196: \ \ \ \ \ \ (ii) for all $\Re \in \mathbf{A}(\Sigma ,\ell )$ and
3197: $s\in \Sigma ^{\ast },$
3198: $$\stackrel{\ell }{\models }rec_{\Re ^{\ast }}(s)\leftrightarrow
3199: (\exists m\geq 0,s_{1},...,s_{m}\in \Sigma ^{\ast
3200: })(s_{1}...s_{m}=s\wedge \wedge _{i=1}^{m}rec_{\Re }(s_{i})).$$
3201:
3202: \smallskip\
3203:
3204: \textbf{Proof.} For (1), (2) and the part from (i) to (ii) of (3),
3205: it is similar to the proof of Proposition 6.6, and here we omit
3206: the details. To show that (ii) implies (i), we assume that
3207: $a,b,c\in L$ and want to construct an $\ell -$valued automaton for
3208: which the validity of (ii) leads to $a\wedge (b\vee c)=(a\wedge
3209: b)\vee (a\wedge c).$ Let $\Re
3210: =<\{q_{1},q_{2},...,q_{6}\},\{q_{1},q_{2},q_{3}\},\{q_{6}\},\delta
3211: >$ in which $\delta (q_{1},\sigma ,q_{4})=\delta (q_{3},\sigma
3212: ,q_{5})=1,$ $\delta (q_{2},\sigma ,q_{6})=a,$ $\delta
3213: (q_{4},\sigma ,q_{6})=b,$ $\delta (q_{5},\sigma ,q_{6})=c,$ and
3214: $\delta $ takes value $0$ for the other arguments. Then $\Re
3215: ^{\ast }$ is visualized as Figure 8.
3216:
3217: \begin{figure}\centering
3218: \includegraphics[width=1\textwidth]{fig_8.eps}\caption{Automaton f} \label{fig 8}
3219: \end{figure}
3220:
3221: We now have
3222: $$a\wedge (b\vee c)=\lceil (\exists m\geq 0,s_{1},...,s_{m}\in
3223: \Sigma ^{\ast })(s_{1}...s_{m}=\sigma ^{3}\wedge \wedge
3224: _{i=1}^{m}rec_{\Re }(s_{i}))\rceil $$
3225: $$=rec_{\Re ^{\ast }}(\sigma ^{3})$$
3226: $$=(a\wedge b)\vee (a\wedge c).\heartsuit$$
3227:
3228: \smallskip\
3229:
3230: From the above proposition, we are able to demonstrate that the
3231: predicate $CReg_\Sigma$ is preserved by the Kleene closure. The
3232: corresponding result for the predicate $Reg_\Sigma$ is not true in
3233: general.
3234:
3235: \smallskip\
3236:
3237: \textbf{Corollary 6.8.} Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
3238: ,0,1>$ be an orthomodular lattice, and let $\rightarrow
3239: =\rightarrow _{3}.$ Then for any $ A\in L^{\Sigma ^{\ast }},$
3240: $$\stackrel{\ell }{\models }CReg_{\Sigma }(A)\longrightarrow
3241: CReg_{\Sigma }(A^{\ast }).$$
3242:
3243: \smallskip\
3244:
3245: \textbf{Proof}. It is similar to the proof of Proposition 6.2. The
3246: point here is to show the following inequality:
3247: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv
3248: rec_{\Re }\rceil \leq \lceil A^{\ast }\equiv rec_{\Re ^{\ast
3249: }}\rceil $$ for any $\Re\in \mathbf{A}(\Sigma, \ell).$ In fact, by
3250: using Lemma 2.11(1) we have
3251: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv
3252: rec_{\Re }\rceil =\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge
3253: \wedge _{s\in \Sigma ^{\ast }}(A(s)\leftrightarrow rec_{\Re
3254: }(s))$$
3255: $$\leq \lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \wedge
3256: _{s\in \Sigma ^{\ast }}(\vee _{s_{1}...s_{m}=s}\wedge
3257: _{i=1}^{m}A(s_{i})\leftrightarrow \vee _{s_{1}...s_{m}=s}\wedge
3258: _{i=1}^{m}rec_{\Re }(s_{i}))$$
3259: $$=\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A^{\ast
3260: }\equiv (rec_{\Re })^{\ast }\rceil .$$
3261:
3262: On the other hand, it follows from Proposition 6.7 that $$\lceil
3263: \gamma (atom(\Re ))\rceil \leq \lceil (rec_{\Re })^{\ast }\equiv
3264: rec_{\Re ^{\ast }}\rceil .$$ Then with Lemma 2.11(3) we obtain
3265: $$\lceil \gamma (atom(\Re )\cup r(A))\rceil \wedge \lceil A\equiv
3266: rec_{\Re }\rceil \leq \lceil \gamma (atom(\Re )\cup r(A))\rceil
3267: \wedge \lceil A^{\ast }\equiv (rec_{\Re })^{\ast }\rceil$$ $$
3268: \wedge \lceil (rec_{\Re })^{\ast }\equiv rec_{\Re ^{\ast }}\rceil
3269: $$
3270: $$\leq \lceil A^{\ast }\equiv rec_{\Re ^{\ast }}\rceil .\heartsuit$$
3271:
3272: \smallskip\
3273:
3274: To conclude this section, we show that both the predicate
3275: $Reg_\Sigma$ and $CReg_\Sigma$ are preserved by the pre-image of a
3276: homomorphism between two languages. But the closure property of an
3277: orthomodular lattice-valued language under homomorphism is left to
3278: be examined in the next section, after the notion of orthomodular
3279: lattice-valued regular expression is proposed. Let $\Sigma $ and
3280: $\Gamma $ be two alphabets of input symbols. Then each mapping
3281: $h:\Sigma \rightarrow \Gamma ^{\ast }$ can be uniquely extended to
3282: a homomorphism $h:\Sigma ^{\ast }\rightarrow \Gamma ^{\ast }$ such
3283: that $ h(\varepsilon )=\varepsilon $ and
3284: $$h(xy)=h(x)h(y)$$ for all $x,y\in \Sigma ^{\ast }.$ Furthermore, we
3285: may define images of $\ell -$valued subsets of $ \Sigma ^{\ast }$
3286: under $h$ and pre-images of $\ell -$valued subsets of $ \Gamma
3287: ^{\ast }$ under $h.$ Recall that for any $A\in L^{\Sigma ^{\ast
3288: }}$ and $B\in L^{\Gamma ^{\ast }},$ $h(A)\in L^{\Gamma ^{\ast }}$
3289: and $h^{-1}(B)\in L^{\Sigma ^{\ast }}$ are given as follows:
3290: $$h(A)(t)=\vee \{A(s):s\in \Sigma ^{\ast }\ {\rm and}\ h(s)=t\}$$
3291: for each $t\in \Gamma ^{\ast },$ and
3292: $$h^{-1}(B)(s)=B(h(s))$$
3293: for each $s\in \Sigma ^{\ast }.$
3294:
3295: Let $\Re =<Q,I,T,\delta >\in \mathbf{A}(\Gamma ,\ell )$ be an
3296: $\ell -$valued automaton over $\Gamma .$ Then the pre-image of
3297: $\Re $ under $h$ is defined to be an $\ell -$valued automaton
3298: $$h^{-1}(\Re )=<Q,I,T,h^{-1}(\delta )>\in \mathbf{A}(\Sigma ,\ell
3299: )$$ over $\Sigma ,$ where for any $p,q\in Q$ and $ \sigma \in
3300: \Sigma ,$
3301: $$h^{-1}(\delta )(p,\sigma ,q)=\delta (p,h(\sigma ),q).$$
3302:
3303: The pre-image of a homomorphism has a very nice compatibility with
3304: the predicates $reg_\Sigma$ and $CReg_\Sigma$, and no
3305: commutativity is needed here.
3306:
3307: \smallskip\
3308:
3309: \textbf{Proposition 6.9.} Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
3310: ,0,1>$ be an orthomodular lattice, let $\rightarrow $ enjoy the
3311: Birkhoff-von Neumann requirement, and let $h:\Sigma \rightarrow
3312: \Gamma ^{\ast }$ be a mapping. Then for any $\Re \in
3313: \mathbf{A}(\Gamma ,\ell )$ and for any $s\in \Sigma ^{\ast },$
3314: $$\stackrel{\ell }{\models} rec_{h^{-1}(\Re )}(s)\longleftrightarrow rec_{\Re
3315: }(h(s)).$$
3316:
3317: \smallskip\
3318:
3319: \textbf{Proof.} Suppose that $s=\sigma _{1}\sigma _{2}...\sigma
3320: _{n}.$ Then
3321: $$\lceil rec_{h^{-1}(\Re )}(s)\rceil =\vee \{I(q_{0})\wedge
3322: T(q_{n})\wedge \wedge _{i=0}^{n-1}h^{-1}(\delta )(q_{i},\sigma
3323: _{i+1},q_{i+1}):q_{0},q_{1},...,q_{n}\in Q\}$$
3324: $$=\vee \{I(q_{0})\wedge T(q_{n})\wedge \wedge _{i=0}^{n-1}\delta
3325: (q_{i},h(\sigma _{i+1}),q_{i+1}):q_{0},q_{1},...,q_{n}\in Q\}$$
3326: $$=\lceil rec_{\Re }(h(\sigma _{1})h(\sigma _{2})...h(\sigma
3327: _{n}))\rceil $$
3328: $$=\lceil rec_{\Re }(h(s))\rceil .\heartsuit$$
3329:
3330: \smallskip\
3331:
3332: \textbf{Corollary 6.10.} Let $\ell =<L,\leq ,\wedge ,\vee ,\perp
3333: ,0,1>$ be an orthomodular lattice, let $\rightarrow $ enjoy the
3334: Birkhoff-von Neumann requirement, and let $h:\Sigma \rightarrow
3335: \Gamma ^{\ast }$ be a mapping. Then for any $B\in L^{\Gamma ^{\ast
3336: }},$
3337: $$\stackrel{\ell }{\models} Reg_{\Gamma }(B)\rightarrow Reg_{\Sigma
3338: }(h^{-1}(B)),$$ and
3339: $$\stackrel{\ell }{\models} CReg_{\Gamma }(B)\rightarrow CReg_{\Sigma
3340: }(h^{-1}(B)),$$
3341:
3342: \smallskip\
3343:
3344: \textbf{Proof.} From the above proposition we have
3345: $$h^{-1}(rec_{\Re })(s)=rec_{\Re }(h(s))=rec_{h^{-1}(\Re )}(s)$$
3346: for all $s\in \Sigma ^{\ast }.$ Then with Lemma 2.12 we obtain
3347: $$\lceil Rec_{\Gamma }(B)\rceil =\vee \{\lceil B\equiv
3348: rec_{\Re }\rceil :\Re \in \mathbf{A}(\Gamma ,\ell )\}$$
3349: $$\leq \vee \{\lceil h^{-1}(B)\equiv h^{-1}(rec_{\Re })\rceil :\Re
3350: \in \mathbf{A}(\Gamma ,\ell )\}$$
3351: $$=\vee \{\lceil h^{-1}(B)\equiv rec_{h^{-1}(\Re )}\rceil :\Re \in \mathbf{A}
3352: (\Gamma ,\ell )\}$$
3353: $$\leq \vee \{\lceil h^{-1}(B)\equiv rec_{\wp }\rceil :\wp \in \mathbf{A}
3354: (\Sigma ,\ell )\}$$
3355: $$=\lceil Reg_{\Sigma }(h^{-1}(B))\rceil .$$
3356:
3357: It is similar for the case of commutative regularity.$\heartsuit$
3358:
3359: \bigskip\
3360:
3361: \textbf{7. Orthomodular Lattice-Valued Regular Expressions}
3362:
3363: \smallskip\
3364:
3365: One of the most interesting results in classical automata theory
3366: is the Kleene theorem which shows the equivalence between finite
3367: automata and regular expressions. The main aim of this section is
3368: to present an orthomodular lattice-valued generalization of the
3369: Kleene theorem. Let $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$ be
3370: an orthomodular lattice, and let $\Sigma$ be an nonempty set of
3371: input symbols. Then the language of $\ell-$valued regular
3372: expressions over $\Sigma$ has the alphabet
3373: $(\Sigma\cup\{\varepsilon,\phi\})\cup(L\cup\{+,\cdot ,*\})$. The
3374: symbols in $\Sigma\cup\{\varepsilon,\phi\}$ will be used to stand
3375: for atomic expressions, and the symbols in $L\cup\{+,\cdot ,*\}$
3376: will be used to denote operators for building up compound
3377: expressions: $*$ and all $\lambda\in L$ are unary operators, and
3378: $+,\cdot$ are binary ones. We use $\alpha,\beta,...$ to act as
3379: meta-symbols for regular expressions and $L(\alpha)$ for the
3380: language denoted by expression $\alpha$. More explicitly,
3381: $L(\alpha)$ will be used to denote an $\ell-$valued subset of
3382: $\Sigma^{*}$; that is, $L(\alpha)\in L^{\Sigma^{*}}$. Orthomodular
3383: lattice-valued regular expressions and the orthomodular
3384: lattice-valued languages denoted by them are formally defined as
3385: follows:
3386:
3387: (i) For each $a\in\Sigma$, $a$ is a regular expression, and
3388: $L(a)={a}$; $\varepsilon$ and $\phi$ are regular expressions, and
3389: $L(\varepsilon)={\varepsilon}$, $L(\phi)=\phi$.
3390:
3391: (ii) If both $\alpha$ and $\beta$ are regular expressions, then
3392: for each $\lambda\in L$, $\lambda\alpha$ is a regular expression,
3393: and $$L(\lambda\alpha)=\lambda L(\alpha);$$ and $\alpha + \beta$,
3394: $\alpha \cdot \beta$ and $\alpha ^{*}$ are all regular
3395: expressions, and
3396: $$L(\alpha +\beta)=L(\alpha)\cup L(\beta),$$ $$L(\alpha \cdot
3397: \beta)=L(\alpha)\cot L(\beta),$$ $$L(\alpha ^{*})=L(\alpha)
3398: ^{*}.$$ It is easy to see that the only difference between
3399: orthomodular lattice-valued regular expressions and the classical
3400: ones is that the additional unary (scalar) operators $\lambda\in
3401: L$ are permitted to occur in the former.
3402:
3403: The central part of the Kleene theorem is a mechanism to transform
3404: a finite automaton into a regular expression. This mechanism has a
3405: straightforward extension in the framework of orthomodular
3406: lattice-valued automata. Let $\Re =<Q,I,T,\delta >\in
3407: \mathbf{A}(\Sigma ,\ell )$ be an $\ell -$valued automaton over
3408: $\Sigma .$ For any $u,v\in Q$ and $X\subseteq Q,$ $\alpha
3409: _{uv}^{X}$ is defined by induction on the cardinality $|X| $ of
3410: $X:$
3411:
3412: (1) $$\alpha _{uv}^{\phi }=\left\{
3413: \begin{array}{c}
3414: \Sigma_{\sigma \in \Sigma }\delta (u,\sigma ,v)\sigma \ {\rm if}\
3415: u\neq
3416: v, \\
3417: \varepsilon +\Sigma_{\sigma \in \Sigma }\delta (u,\sigma ,v)\sigma
3418: \ {\rm if}\ u=v.
3419: \end{array}
3420: \right. $$
3421:
3422: (2) if $X\neq \phi ,$ then we choose any $q\in X$ and define
3423: $$\alpha _{uv}^{X}=\alpha _{uv}^{X-\{q\}}+\alpha
3424: _{uq}^{X-\{q\}}\cdot (\alpha _{qq}^{X-\{q\}})^{\ast }\cdot \alpha
3425: _{qv}^{X-\{q\}}.$$ Then the $\ell -$valued regular expression
3426: $$k(\Re )=\Sigma_{u,v\in Q}(I(u)\wedge T(v))\alpha
3427: _{uv}^{Q}$$ is called a Kleene representation of $\Re .$
3428:
3429: The following theorem describes properly the relationship between
3430: the language recognized by an orthomodular lattice-valued
3431: automaton and the language expressed by its Kleene representation.
3432:
3433: \smallskip\
3434:
3435: \textbf{Theorem 7.1. } Let $\ell =<L,\leq ,\wedge ,\vee ,\bot
3436: ,0,1>$ be an orthomodular lattice, and let $\rightarrow$ satisfy
3437: the Birkhoff-von Neumann requirement.
3438:
3439: (1) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and $s\in \Sigma
3440: ^{\ast },$ if $k(\Re)$ is a Kleene representation of $\Re$, then
3441: $$\stackrel{\ell }{\models} rec_{\Re }(s)\longrightarrow s\in L(k(\Re
3442: )).$$
3443:
3444: (2) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and $s\in \Sigma
3445: ^{\ast },$ and for any Kleene representation $k(\Re)$ of $\Re$,
3446: $$\stackrel{\ell }{\models} \gamma(atom(\Re))\wedge s\in L(k(\Re))\longrightarrow rec_\Re
3447: (s),$$ and especially if $\longrightarrow =\longrightarrow_3$,
3448: then $$\stackrel{\ell }{\models} \gamma(atom(\Re))\longrightarrow
3449: (rec_\Re (s)\longleftrightarrow s\in L(k(\Re))).$$
3450:
3451: (3) The following two statements are equivalent:
3452:
3453: \ \ \ \ \ \ (i) $\ell $ is a Boolean algebra.
3454:
3455: \ \ \ \ \ \ (ii) For any $\Re \in \mathbf{A}(\Sigma ,\ell )$ and
3456: $s\in \Sigma ^{\ast },$ and for any Kleene representation $k(\Re)$
3457: of $\Re$,
3458: $$\stackrel{\ell }{\models} rec_{\Re }(s)\longleftrightarrow s\in L(k(\Re
3459: )).$$
3460:
3461: \smallskip\
3462:
3463: \textbf{Proof.} We prove (1) and (2) together. To this end, we
3464: have to demonstrate that for any $u,v\in Q,$ $X\subseteq Q$ and
3465: $s\in \Sigma ^{\ast },$ $$(a)\ \vee \{\lceil path_{\Re
3466: }(c)\rceil:c\in T(Q,\Sigma ),b(c)=u,e(c)=v,M(c)\subseteq
3467: X,lb(c)=s\}\leq L(\alpha _{uv}^{X})(s),$$
3468: $$(b)\ \lceil\gamma(atom(\Re))\rceil\wedge L(\alpha _{uv}^{X})(s)\leq\ \ \ \ \ \
3469: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3470: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3471: $$
3472: $$\ \ \ \ \ \ \ \ \vee \{\lceil path_{\Re }(c)\rceil:c\in T(Q,\Sigma ),b(c)=u,e(c)=v,M(c)\subseteq
3473: X,lb(c)=s\},$$ where $M(c)$ stands for the set of states along $c$
3474: except $u$ and $v;$ more exactly, $M(c)=\{q_{1},...,q_{k-1}\}$ if
3475: $c=u\sigma _{1}q_{1}...q_{k-1}\sigma _{k}v.$ This claim may be
3476: proved by induction on $ |X|.$ For the case of $X=\phi ,$ it is
3477: easy. We now suppose that $q\in X\neq \phi $ and $$\alpha
3478: _{uv}^{X}=\alpha _{uv}^{X-\{q\}}+[\alpha _{uq}^{X-\{q\}}(\alpha
3479: _{qq}^{X-\{q\}})^{\ast }]\alpha _{qv}^{X-\{q\}}.$$ We first show
3480: that (a) is valid in this case. From the induction hypothesis we
3481: have $$(c)\ \vee \{\lceil path_{\Re }(c)\rceil :c\in T(Q,\Sigma
3482: ),b(c)=e(c)=q,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3483: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3484: $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3485: M(c)\subseteq X-\{q\}, lb(c)=s\}\leq L(\alpha
3486: _{qq}^{X-\{q\}})(s)$$ for each $s\in \Sigma ^{\ast }.$ Then we
3487: assert that for all $s\in \Sigma ^{\ast},$
3488: $$(d)\ \vee \{\lceil path_{\Re }(c)\rceil :c\in
3489: T(Q,\Sigma ),b(c)=e(c)=q,M(c)\subseteq X,lb(c)=s\}
3490: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3491: $$\leq L((\alpha
3492: _{qq}^{X-\{q\}})^{\ast })(s).$$ In fact, for any $c\in T(Q,\Sigma
3493: )$ if $b(c)=e(c)=q,M(c)\subseteq X$ and $ lb(c)=s,$ we write
3494: $c_{i}$ for the substring of $c$ beginning with the $i$th $q $ and
3495: ending at the $(i+1)$th $q.$ If the number of occurrences of $q$
3496: in $c$ is $m+1,$ then
3497: $$\lceil path_{\Re }(c)\rceil =\wedge _{i=1}^{m}\lceil path_{\Re }(c_{i})\rceil .$$
3498: Furthermore, by using (c) and noting that
3499: $s=lb(c_{1})...lb(c_{m})$ we obtain
3500: $$\lceil path_{\Re }(c)\rceil =\wedge _{i=1}^{m}L(\alpha
3501: _{qq}^{X-\{q\}})(lb(c_{i}))\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3502: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3503: \ \ \ \ \ \ \ $$
3504: $$\leq\vee\{\wedge_{i=1}^{n} L(\alpha_{qq}^{X-\{q\}})(s_i):n\geq
3505: 0, s_1,...,s_n\in \Sigma^{*},s=s_1...s_n\}$$
3506: $$=(L(\alpha _{qq}^{X-\{q\}}))^{\ast }(s)$$
3507: $$= L((\alpha _{qq}^{X-\{q\}})^{\ast })(s).$$
3508: Let $c$ range over $\{c\in T(Q,\Sigma ):b(c)=e(c)=q,M(c)\subseteq
3509: X,lb(c)=s\}.$ Then (d) is proved.
3510:
3511: Furthermore, from the induction hypothesis and (d) we have
3512: $$([L(\alpha _{uq}^{X-\{q\}})L((\alpha _{qq}^{X-\{q\}})^{\ast
3513: })]L(\alpha _{qv}^{X-\{q\}}))(s)=\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3514: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3515: \ \ $$
3516: $$\vee \{[L(\alpha_{uq}^{X-\{q\}})L((\alpha_{qq}^{X-\{q\}})^{\ast})](x)
3517: \wedge L(\alpha_{qv}^{X-\{q\}})(y):s=xy\} $$
3518: $$=\vee \{\vee\{L(\alpha_{uq}^{X-\{q\}})(x_1)\wedge
3519: L((\alpha_{qq}^{X-\{q\}})^{\ast})(x_2):x=x_1 x_2\} \wedge
3520: L(\alpha_{qv}^{X-\{q\}})(y):s=xy\}\ \ $$
3521: $$\geq \vee \{L(\alpha_{uq}^{X-\{q\}})(x_1)\wedge
3522: L((\alpha_{qq}^{X-\{q\}})^{\ast})(x_2) \wedge
3523: L(\alpha_{qv}^{X-\{q\}})(y):s=x_1x_2y\}\ \ \ \ \ \ \ \ \ \ \ \ \ \
3524: \ \ \ $$
3525: $$\geq \vee \{\lceil path_{\Re }(c_{1})\rceil \wedge \lceil path_{\Re }(c_{2})\rceil \wedge \lceil path_{\Re
3526: }(c_{3})\rceil : c_{1},c_{2},c_{3}\in T(Q,\Sigma),\ \ \ \ \ \ \ \
3527: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3528: $$\ \ \ \ b(c_{1})=u, e(c_{1})=b(c_{2})=e(c_{2})=b(c_{3})=q, e(c_{3})=v,s=lb(c_{1})lb(c_{2})lb(c_{3})\}$$
3529: $$=\vee \{\lceil path_{\Re }(c)\rceil :c\in T(Q,\Sigma ),b(c)=u,e(c)=v,q\in
3530: M(c)\}.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3531: \ \ $$ This yields further
3532: $$L(\alpha _{uv}^{X})(s)=L(\alpha _{uv}^{X-\{q\}})(s)\vee
3533: ([L(\alpha _{uq}^{X-\{q\}})L((\alpha _{qq}^{X-\{q\}})^{\ast
3534: }]L(\alpha _{qv}^{X-\{q\}}))(s)$$
3535: $$\geq \ {\rm the\ left-hand\ side\ of\ (a)}.$$
3536:
3537: We now turn to consider (b). The induction hypothesis gives
3538: $$(e)\ \lceil \gamma(atom(\Re))\rceil \wedge L(\alpha_{uv}^{X-\{q\}})(s)\leq \{\lceil path_\Re (c)\rceil :c\in
3539: T(Q,\Sigma),\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3540: \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3541: $$b(c)=u, e(c)=v, M(c)\subseteq X-\{q\}, lb(c)=s \}.$$ For any
3542: $n\geq 0$ and $s_{1},...,s_{n}\in \Sigma ^{\ast }$ with $
3543: s=s_{1}...s_{n},$ from (e) we can apply Lemmas 2.5 and 2.6 to
3544: obtain
3545: $$\lceil \gamma(atom(\Re))\rceil \wedge \wedge
3546: _{i=1}^{n}L(\alpha _{qq}^{X-\{q\}})(s_{i})=\lceil
3547: \gamma(atom(\Re))\rceil \wedge \wedge_{i=1}^{n}[\lceil
3548: \gamma(atom(\Re))\rceil \wedge L(\alpha_{qq}^{X-\{q\}})(s_i)]$$
3549: $$\leq \lceil \gamma(atom(\Re))\rceil \wedge \wedge _{i=1}^{n}\vee \{\lceil path_{\Re
3550: }(c_{i})\rceil :c_{i}\in T(Q,\Sigma ),\ \ \ \ \ \ \ \ \ \ \ \ \ \
3551: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3552: \
3553: $$
3554: $$b(c_{i})=e(c_{i})=q,M(c_{i})\subseteq X-\{q\},lb(c_{i})=s_{i}\}$$
3555: $$\leq \vee \{\wedge _{i=1}^{n}\lceil path_{\Re }(c_{i})\rceil :c_{i}\in T(Q,\Sigma
3556: ),b(c_{i})=e(c_{i})=q, M(c_{i})\subseteq X-\{q\},\ \ \ \ \ \ \ \ \
3557: \ \ \ \ \ \ \ \ \ \ $$
3558: $$lb(c_{i})=s_{i}\ {\rm for\ each}\ i=1,2,...,n\}$$
3559: $$\leq \vee \{\lceil path_{\Re }(\overline{c_{1}...c_{n}})\rceil :c_{i}\in T(Q,\Sigma
3560: ),b(c_{i})=e(c_{i})=q, M(c_{i})\subseteq X-\{q\},\ \ \ \ \ \ \ \ \
3561: \ \ \ \ \ \ \ \ \ \ $$
3562: $$lb(c_{i})=s_{i}\ {\rm for\ each}\ i=1,2,...,n\},$$ where
3563: $\overline{c_{1}...c_{n}}=c_{1}c_{2}^{\prime }...c_{n}^{\prime },$
3564: $c_{i}^{\prime }$ is the resulting string after removing the first
3565: $q$ in $c_{i}$ for each $i\geq 2.$ Note that
3566: $lb(\overline{c_{1}...c_{n}} )=s_{1}...s_{n}=s$ whenever
3567: $lb(c_{i})=s_{i}$ $(i=1,2,...,n).$ We write
3568: $$\lambda=\vee \{\lceil path_\Re (c)\rceil :c\in T(Q,\Sigma), b(c)=e(c)=q, M(c)\subseteq X,
3569: lb(c)=s\}.$$ Then it holds that
3570: $$\lceil \gamma(atom(\Re))\rceil \wedge \wedge _{i=1}^{n}L(\alpha _{qq}^{X-\{q\}})(s_{i})\leq\lambda.$$
3571: Moreover, note that $\lceil \gamma(atom(\Re))\rceil ,
3572: L(\alpha_{qq}^{X-\{q\}})(s_i)\in [atom(\Re)].$ It follows that
3573: $$\lceil \gamma(atom(\Re))\rceil \wedge L((\alpha_{qq}^{X-\{q\}})^{\ast})(s)=\lceil \gamma(atom(\Re))\rceil \wedge
3574: \lceil \gamma(atom(\Re))\rceil \wedge\ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3575: \ \ \ \ \ \ \ \ \ \ \ \ $$
3576: $$\vee\{\wedge_{i=1}^{n}L(\alpha_{qq}^{X-\{q\}})(s_i):n\geq 0,
3577: s=s_1...s_n \}$$
3578: $$\leq \vee\{\lceil \gamma(atom(\Re))\rceil \wedge \wedge_{i=1}^{n}L(\alpha_{qq}^{X-\{q\}})(s_i):n\geq 0,
3579: s=s_1...s_n \}\leq \lambda.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3580: $$
3581: This enables us to obtain
3582: $$\lceil \gamma (atom(\Re ))\rceil \wedge \lbrack L(\alpha
3583: _{uq}^{X-\{q\}})L((\alpha _{qq}^{X-\{q\}})^{\ast })](x)\ \ \ \ \ \
3584: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3585: \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3586: $$=\lceil \gamma
3587: (atom(\Re ))\rceil \wedge \lceil \gamma (atom(\Re ))\rceil \wedge
3588: \vee \{L(\alpha _{uq}^{X-\{q\}})(x_{1})\wedge L((\alpha
3589: _{qq}^{X-\{q\}})^{\ast })(x_{2}):x=x_{1}x_{2}\}\ \ \ $$
3590: $$\leq \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge
3591: L(\alpha _{uq}^{X-\{q\}})(x_{1})\wedge L((\alpha
3592: _{qq}^{X-\{q\}})^{\ast })(x_{2}):x=x_{1}x_{2}\}\ \ \ \ \ \ \ \ \ \
3593: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3594: $$=\vee \{\lceil \gamma
3595: (atom(\Re ))\rceil \wedge [\lceil \gamma (atom(\Re ))\rceil \wedge
3596: L(\alpha _{uq}^{X-\{q\}})(x_{1})]\wedge\ \ \ \ \ \ \ \ \ \ \ \ \ \
3597: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3598: \ \ $$
3599: $$[\lceil \gamma (atom(\Re ))\rceil \wedge
3600: L((\alpha _{qq}^{X-\{q\}})^{\ast })(x_{2})]:x=x_{1}x_{2}\}$$
3601: $$\leq
3602: \vee \{\lceil \gamma (atom(\Re ))\rceil \wedge [ \vee \{\lceil
3603: path_{\Re }(c_{1})\rceil :c_{1}\in T(Q,\Sigma
3604: ),b(c_{1})=u,e(c_{1})=q,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3605: \ \ \ \ \ \ $$
3606: $$M(c_{1})\subseteq X-\{q\},lb(c_{1})=x_{1}\}]
3607: \wedge \lbrack \vee \{\lceil path_{\Re }(c_{2})\rceil :c_{2}\in
3608: T(Q,\Sigma ),b(c_{2})=$$ $$e(c_{2})=q,M(c_{2})\subseteq
3609: X,lb(c_{2})=x_{2}\}]:x=x_{1}x_{2}\}$$
3610: $$\leq \vee \{\lceil path_{\Re
3611: }(c_{1})\rceil \wedge \lceil path_{\Re }(c_{2})\rceil
3612: :c_{1},c_{2}\in T(Q,\Sigma
3613: ),b(c_{1})=u,e(c_{1})=b(c_{2})=e(c_{2})=q,$$
3614: $$M(c_{1})\subseteq X-\{q\}, M(c_{2})\subseteq
3615: X,x=lb(c_{1})lb(c_{2})\}.$$ Furthermore, we can derive in a
3616: similar way that
3617: $$\lceil \gamma (atom(\Re
3618: ))\rceil \wedge ([L(\alpha _{uq}^{X-\{q\}})L((\alpha
3619: _{qq}^{X-\{q\}})^{\ast })]L(\alpha _{qv}^{X-\{q\}}))(s)\ \ \ \ \ \
3620: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3621: $$\leq \vee \{\lceil path_{\Re }(c_{1})\rceil \wedge \lceil path_{\Re }(c_{2})\rceil \wedge \lceil path_{\Re
3622: }(c_{3})\rceil :c_{1},c_{2},c_{3}\in T(Q,\Sigma ),b(c_{1})=u,\ \ \
3623: \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3624: $$e(c_{1})=b(c_{2})=e(c_{2})=b(c_{3})=q,e(c_{3})=v,s=lb(c_{1})lb(c_{2})lb(c_{3})\}
3625: $$
3626: $$=\vee \{\lceil path_{\Re }(c)\rceil :c\in T(Q,\Sigma ),b(c)=u,e(c)=v,q\in
3627: M(c),s=lb(c)\}.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3628: Consequently, it holds that
3629: $$\lceil \gamma (atom(\Re ))\rceil \wedge L(\alpha _{uv}^{X})(s)=\lceil \gamma (atom(\Re
3630: ))\rceil \wedge \{L(\alpha _{uv}^{X-\{q\}})(s)\vee\ \ \ \ \ \ \ \
3631: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3632: $$\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ ([L(\alpha _{uq}^{X-\{q\}})L((\alpha _{qq}^{X-\{q\}})^{\ast
3633: })]L(\alpha _{qv}^{X-\{q\}}))(s)\}$$
3634: $$\leq [\lceil \gamma (atom(\Re ))\rceil \wedge L(\alpha
3635: _{uv}^{X-\{q\}})(s)]\vee \{\lceil \gamma (atom(\Re ))\rceil \wedge
3636: ([L(\alpha _{uq}^{X-\{q\}})L((\alpha _{qq}^{X-\{q\}})^{\ast
3637: })]L(\alpha _{qv}^{X-\{q\}}))(s)\}$$
3638: $$\leq\ {\rm the\ right-hand\ side\ of\ (b)}.\ \ \ \ \ \
3639: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3640: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3641: $$
3642:
3643: After proving (a), we can assert that
3644: $$\lceil s\in L(k(\Re ))\rceil =\vee
3645: _{u,v\in Q}[I(u)\wedge T(v)\wedge L(\alpha _{uv}^{Q})(s)]\ \ \ \ \
3646: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3647: \ \ \ \ \ $$
3648: $$\geq \vee _{u,v\in Q}[(I(u)\wedge T(v))\wedge \vee \{\lceil path_{\Re
3649: }(c)\rceil :c\in T(Q,\Sigma ),b(c)=u,e(c)=v,lb(c)=s\}]$$
3650: $$\geq \vee _{u,v\in Q}\vee \{I(u)\wedge T(v)\wedge \lceil path_{\Re }(c)\rceil :c\in
3651: T(Q,\Sigma ),b(c)=u,e(c)=v,lb(c)=s\}\ \ $$
3652: $$=\lceil rec_{\Re }(s)\rceil .\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3653: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3654: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3655:
3656: By using (b) and Lemmas 2.5 and 2.6, we have
3657: $$\lceil \gamma (atom(\Re))\rceil \wedge \lceil s\in L(k(\Re ))\rceil =\lceil \gamma (atom(\Re))\rceil \wedge \vee
3658: _{u,v\in Q}[I(u)\wedge T(v)\wedge L(\alpha _{uv}^{Q})(s)]$$
3659: $$\leq \vee
3660: _{u,v\in Q}[I(u)\wedge T(v)\wedge \lceil \gamma (atom(\Re))\rceil
3661: \wedge L(\alpha _{uv}^{Q})(s)]\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3662: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3663: $$\leq \vee _{u,v\in Q}[(I(u)\wedge T(v))\wedge \lceil \gamma (atom(\Re))\rceil\wedge \vee \{\lceil path_{\Re
3664: }(c)\rceil :c\in T(Q,\Sigma ),\ \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3665: $$b(c)=u,e(c)=v,lb(c)=s\}]$$
3666: $$\leq \vee _{u,v\in Q}\vee \{I(u)\wedge T(v)\wedge \lceil path_{\Re }(c)\rceil :c\in
3667: T(Q,\Sigma ),b(c)=u,e(c)=v,lb(c)=s\}$$
3668: $$=\lceil rec_{\Re }(s)\rceil .\ \ \ \ \ \ \
3669: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3670: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3671: \ \ \ \ \ \ \ \ \ \ \ \ \ $$
3672:
3673: Thus, (1) and (2) are proved, and the part that (i) implies (ii)
3674: of (3) is a simple corollary of (2). We now turn to prove that
3675: (ii) implies (i). For any $a,b,c\in L,$ we consider the $\ell
3676: -$valued automaton
3677: $$\Re =<\{u,v\},\delta ,u_{a},\{u,v\}>,$$ where $\delta (u,\sigma
3678: ,u)=b,$ $\delta (u,\sigma ,v)=c,$ and $\delta $ takes value $0$
3679: for other cases (see Figure 9). Then
3680: $$\lceil rec_{\Re }(\sigma )\rceil =\vee \{I(q_{0})\wedge
3681: T(q_{1})\wedge \delta (q_{0},\sigma ,q_{1}):q_{0},q_{1}\in Q\}$$
3682: $$=(a\wedge b)\vee (a\wedge c).$$
3683:
3684: \begin{figure}\centering
3685: \includegraphics{fig_9.eps}\caption{Automaton g} \label{fig 9}
3686: \end{figure}
3687:
3688: On the other hand, we have
3689: $$\left\{
3690: \begin{array}{c}
3691: \alpha _{uu}^{\phi }=\varepsilon +b\sigma , \\
3692: \alpha _{uv}^{\phi }=c\sigma , \\
3693: \alpha _{vv}^{\phi }=\varepsilon , \\
3694: \alpha _{vu}^{\phi }=\phi .
3695: \end{array}
3696: \right. $$ Therefore,
3697: $$\alpha _{uu}^{\{v\}}=\alpha
3698: _{uu}^{\phi }+[\alpha _{uv}^{\phi }(\alpha _{vv}^{\phi })^{\ast
3699: }]\alpha _{vu}^{\phi }$$
3700: $$=(\varepsilon +b\sigma )+[c\sigma
3701: (\varepsilon )^{\ast }]\phi $$
3702: $$=\varepsilon +b\sigma ,$$
3703: $$\alpha _{uv}^{\{v\}}=\alpha _{uv}^{\phi }+[\alpha _{uv}^{\phi
3704: }(\alpha _{vv}^{\phi })^{\ast }]\alpha _{vv}^{\phi }$$
3705: $$=c\sigma +[c\sigma (\varepsilon )^{\ast }]\varepsilon $$
3706: $$=c\sigma ,$$
3707: and
3708: $$\alpha _{uv}^{\{u,v\}}=\alpha _{uu}^{\{v\}}+[\alpha _{uu
3709: }^{\{v\}}(\alpha _{u}^{\{v\}})^{\ast }]\alpha _{uv}^{\{v\}}$$
3710: $$=\varepsilon +b\sigma +[(\varepsilon +b\sigma )(\varepsilon
3711: +b\sigma )^{\ast }](c\sigma ).$$
3712:
3713: From the assumption (ii) we know that
3714: $$(a\wedge b)\vee (a\wedge
3715: c)=\lceil rec_{\Re }(\sigma )\rceil $$
3716: $$=L(k(\Re ))(\sigma )$$
3717: $$=[L(a\alpha _{u}^{\{u,v\}})\cup L(a\alpha
3718: _{uv}^{\{u,v\}})](\sigma )$$
3719: $$\geq L(a\alpha _{u}^{\{u,v\}})(\sigma )$$
3720: $$=a\wedge L(\alpha _{u}^{\{u,v\}})(\sigma )$$
3721: $$=a\wedge L(\varepsilon +b\sigma +[(\varepsilon +b\sigma
3722: )(\varepsilon +b\sigma )^{\ast }](c\sigma ))(\sigma )$$
3723: $$\geq
3724: a\wedge (b\vee c).$$ This completes the proof.$\heartsuit$
3725:
3726: \smallskip\
3727:
3728: \textbf{Corollary 7.2.} Let $\ell =<L,\leq ,\wedge ,\vee ,\bot
3729: ,0,1>$ be an orthomodular lattice, and let
3730: $\rightarrow=\rightarrow_3$. Then for any $A\in L^{\Sigma^{*}}$,
3731: $$\stackrel{\ell}{\models} CReg_{\Sigma}(A)\rightarrow (\exists\ {\rm regular\ expression}\ \alpha)(A\equiv L(\alpha)).$$
3732:
3733: \smallskip\
3734:
3735: \textbf{Proof.} It can be derived from Theorem 7.1 in a way
3736: similar to the proof of Corollary 4.2.$\heartsuit$
3737:
3738: \smallskip\
3739:
3740: We now turn to consider homomorphisms of $\ell-$valued regular
3741: expressions. Let $\Sigma$ and $\Gamma$ be two alphabet, and let
3742: $h:\Sigma \rightarrow \Gamma ^{\ast }$ be a mapping. Then it can
3743: be uniquely extended to a mapping, denoted still by $h,$ from
3744: $\ell -$valued regular expressions over $\Sigma $ into $\ell
3745: -$valued regular expressions over $\Gamma .$ For any $\ell
3746: -$valued regular expression $\alpha $ over $ \Sigma ,$ $h(\alpha
3747: )$ is defined to be the $\ell -$valued regular expression over
3748: $\Gamma $ obtained by replacing each letter $\sigma \in \Sigma $
3749: appearing in $\alpha $ with the string $h(\sigma )\in \Gamma
3750: ^{\ast }.$ Formally, $h(\alpha )$ is defined by induction on the
3751: length of $\alpha : $
3752: $$h(\varepsilon )=\varepsilon ,$$
3753: $$h(\phi )=\phi ,$$
3754: $$h(\sigma )\ {\rm is\ already\ given\ for\ each}\ \sigma \in \Sigma
3755: ,$$
3756: $$h(\lambda \alpha )=\lambda h(\alpha ),$$
3757: $$h(\alpha _{1}+\alpha _{2})=h(\alpha _{1})+h(\alpha _{2}),$$
3758: $$h(\alpha _{1}\cdot\alpha _{2})=h(\alpha _{1})\cdot h(\alpha _{2}),$$
3759: $$h(\alpha ^{\ast })=(h(\alpha ))^{\ast }.$$
3760:
3761: \smallskip\
3762:
3763: For each $\ell-$valued regular expression $\alpha$ over $\Sigma$,
3764: we write $\Lambda (\alpha )$ for the set of scalar values $\lambda
3765: \in L$ occurring in $\alpha .$ Indeed, $\Lambda (\alpha )$ may be
3766: formally defined by induction on the length of $\alpha $ as
3767: follows:
3768: $$\Lambda (\varepsilon )=\Lambda (\phi )=\Lambda (\sigma )=\phi\ {\rm for\ every}\
3769: \sigma \in \Sigma ,$$
3770: $$\Lambda (\lambda \alpha )=\{\lambda \}\cup
3771: \Lambda (\alpha ),$$
3772: $$\Lambda (\alpha _{1}+\alpha _{2})=\Lambda (\alpha _{1}\cdot \alpha
3773: _{2})=\Lambda (\alpha _{1})\cup \Lambda (\alpha _{2}),$$
3774: $$\Lambda
3775: (\alpha ^{\ast })=\Lambda (\alpha ).$$ It is easy to see that
3776: $\Lambda (\alpha )$ is a finite subset of $L.$ Moreover, we write
3777: $$\Delta(\alpha)=\{\mathbf{a}:a\in \Lambda (\alpha)\}$$ for the
3778: set of (constant) propositions in our logical language
3779: corresponding to the elements in $\Lambda (\alpha)$.
3780:
3781: The following two lemmas are very useful when we are dealing with
3782: orthomodular lattice-valued expressions, they evaluate the range
3783: of language generated by an orthomodular lattice-valued regular
3784: expression. In particular, it will be shown in Lemma 7.4 that this
3785: range is a finite set whenever the lattice $\ell$ of truth values
3786: is a Boolean algebra.
3787:
3788: \smallskip\
3789:
3790: \textbf{Lemma 7.3.} Let $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$
3791: be an orthomodular lattice. Then for any $\ell -$valued regular
3792: expression $\alpha ,$ $ \{L(\alpha )(s):s\in \Sigma ^{\ast
3793: }\}\subseteq [\Lambda (\alpha )],$ where $[A]$ denotes the
3794: subalgebra of $\ell$ generated by $A$ for any $A\subseteq L.$
3795:
3796: \smallskip\
3797:
3798: \textbf{Proof.} We use an induction argument on the length of
3799: $\alpha .$ For simplicity, we only consider the following two
3800: cases, and the other cases are easy or similar.
3801:
3802: (1) From the induction hypothesis we know that
3803: $$L(\lambda .\alpha
3804: )(s)=\lambda \wedge L(\alpha )(s)\in \overline{\{\lambda \}\cup
3805: \Lambda (\alpha )}=\overline{\Lambda (\lambda .\alpha )}$$ for
3806: each $s\in \Sigma ^{\ast }.$
3807:
3808: (2) Let $s\in \Sigma ^{\ast }.$ For any $s_{1},...,s_{n}\in \Sigma
3809: ^{\ast }$ with $s_{1}...s_{n}=s,$ we suppose that
3810: $s_{i_{1}},...,s_{i_{m}}\neq \varepsilon $ and $s_{i}=\varepsilon
3811: $ for every $i\in \{1,...,n\}-\{i_{1},...,i_{m}\}.$ Then
3812: $s_{i_{1}}...s_{i_{m}}=s$ and
3813: $$L(\alpha )(s_{1})\wedge ...\wedge
3814: L(\alpha )(s_{n})=\left\{
3815: \begin{array}{c}
3816: L(\alpha )(s_{i_{1}})\wedge ...\wedge L(\alpha )(s_{i_{m}})\ {\rm
3817: if }\ m=n,
3818: \\
3819: L(\alpha )(s_{i_{1}})\wedge ...\wedge L(\alpha )(s_{i_{m}})\wedge
3820: L(\alpha )(\varepsilon )\ {\rm if }\ m<n.
3821: \end{array}
3822: \right. $$ Furthermore, we note that $$\{(s_{1},...,s_{n}):n\geq
3823: 0,s_{1},...,s_{n}\in \Sigma ^{\ast }-\{\varepsilon \}\ {\rm and}\
3824: s_{1}...s_{n}=s\}$$ is finite. Therefore, $$\{L(\alpha
3825: )(s_{1})\wedge ...\wedge L(\alpha )(s_{n}):s_{1}...s_{n}=s\}$$ is
3826: also finite, and with the induction hypothesis we have
3827: $$L(\alpha ^{\ast })(s)=\vee \{L(\alpha )(s_{1})\wedge ...\wedge
3828: L(\alpha )(s_{n}):s_{1}...s_{n}=s\}\in \overline{\Lambda (\alpha
3829: )}.\heartsuit$$
3830:
3831: \smallskip\
3832:
3833: \textbf{Lemma 7.4.} If $\ell =<L,\leq ,\wedge ,\vee ,\bot ,0,1>$
3834: is a Boolean algebra, then for any $\ell -$valued regular
3835: expression $\alpha ,$ $ \{L(\alpha )(s):s\in \Sigma ^{\ast }\}$ is
3836: a finite set.
3837:
3838: \smallskip\
3839:
3840: \textbf{Proof. }From Lemma 7.3 and the distributivity of $\wedge $
3841: over $\vee $ we know that for any $s\in \Sigma ^{\ast },$ there
3842: are $\lambda _{ij_{i}}\in \Lambda (\alpha )$
3843: $(i=1,...,m;j_{i}=1,...,n_{i})$ such that
3844: $$L(\alpha )(s)=\vee
3845: _{i=1}^{m}(\wedge _{j_{i}=1}^{n_{i}}\lambda _{ij_{i}}).$$ Since
3846: $\Lambda (\alpha )$ is finite, both $$\Lambda (\alpha )^{(\wedge
3847: )}=\{\lambda _{1}\wedge ...\wedge \lambda _{n}:n\geq 0,\lambda
3848: _{1},...,\lambda \in \Lambda (\alpha )\}$$ and $$\Lambda (\alpha
3849: )^{(\wedge )(\vee )}=\{\vee M:M\subseteq \Lambda (\alpha
3850: )^{(\wedge )}\}$$ are also finite. Therefore, $$\Lambda (\alpha
3851: )^{(\wedge )(\vee )}\supseteq \{L(\alpha )(s):s\in \Sigma ^{\ast
3852: }\}$$ is a finite set.$\heartsuit$
3853:
3854: \smallskip\
3855:
3856: The following proposition shows that a homomorphism preserves the
3857: language generated by an orthomodular lattice-valued regular
3858: expression under the condition that all elements in the range of
3859: the expression under consideration are commutative.
3860:
3861: \smallskip\
3862:
3863: \textbf{Proposition 7.5.} Let $\ell =<L,\leq ,\wedge ,\vee ,\bot
3864: ,0,1>$ be an orthomodular lattice and $\rightarrow$ fulfil the
3865: Birkhoff-von Neumann requirement, and let $\Sigma$ and $\Gamma$ be
3866: two alphabets.
3867:
3868: (1) For any mapping $h:\Sigma\rightarrow \Gamma ^{\ast }$, and for
3869: any $\ell -$valued regular expression $\alpha $ over $\Sigma ,$
3870: $$\stackrel{\ell}{\models} h(L(\alpha ))\subseteq L(h(\alpha )).$$
3871:
3872: (2) For any mapping $h:\Sigma\rightarrow \Gamma ^{\ast }$, for any
3873: $\ell -$valued regular expression $\alpha $ over $\Sigma ,$ and
3874: for any $t\in \Gamma^{*}$,
3875: $$\stackrel{\ell}{\models}\gamma(\Delta(\alpha))\wedge t\in L(h(\alpha))\rightarrow
3876: t\in h(L(\alpha)),$$ and if $\rightarrow=\rightarrow_3$ then
3877: $$\stackrel{\ell}{\models}\gamma(\Delta(\alpha))\rightarrow
3878: L(h(\alpha))\equiv h(L(\alpha)).$$
3879:
3880: (3) The following two statements are equivalent:
3881:
3882: \ \ \ \ \ \ (i) $\ell $ is a Boolean algebra.
3883:
3884: \ \ \ \ \ \ (ii) for any mapping $h:\Sigma\rightarrow \Gamma
3885: ^{\ast }$, and for any $\ell-$valued regular expression $ \alpha $
3886: over $\Sigma ,$
3887: $$\stackrel{\ell}{\models} h(L(\alpha ))\equiv L(h(\alpha )).$$
3888:
3889: \smallskip\
3890:
3891: \textbf{Proof.} We only prove (2) and (3), and (1) can be observed
3892: from the proof of (2). The part that (i) implies (ii) of (3) may
3893: be derived from (2); and it can also be proved directly by using
3894: Lemma 7.4.
3895:
3896: Our first aim is to prove that $$\lceil
3897: \gamma(\Delta(\alpha))\rceil\wedge L(h(\alpha ))(t)=h(L(\alpha
3898: ))(t)$$ for any $t\in \Gamma ^{\ast }$ and for any $\ell -$valued
3899: regular expression $\alpha $ over $ \Sigma .$ We proceed by
3900: induction on the length of $\alpha .$
3901:
3902: (a) It is obvious for the case of $\alpha =\varepsilon $ or $\phi
3903: ,$ or $ \alpha \in \Sigma .$
3904:
3905: (b) With the definitions of $h(\cdot )$ and $L(\cdot )$ and the
3906: induction hypothesis we derive that
3907: $$L(h(\lambda .\alpha ))(t)=L(\lambda .h(\alpha ))(t)$$
3908: $$=\lambda \wedge L(h(\alpha ))(t)$$
3909: $$=\lambda \wedge h(L(\alpha ))(t)$$
3910: $$=\lambda \wedge \vee \{L(\alpha )(s):s\in \Sigma ^{\ast }\ {\rm
3911: and}\ h(s)=t\}.$$ Then from Lemmas 2.5, 2.6 and 7.3, it follows
3912: that
3913: $$\lceil \gamma(\Delta(\alpha))\rceil\wedge L(h(\lambda
3914: .\alpha ))(t)\leq \vee \{\lambda \wedge L(\alpha )(s):s\in \Sigma
3915: ^{\ast }\ {\rm and}\ h(s)=t\}$$
3916: $$=\vee \{L(\lambda .\alpha )(s):s\in
3917: \Sigma ^{\ast }\ {\rm and}\ h(s)=t\}$$
3918: $$=h(L(\lambda .\alpha
3919: ))(t).$$
3920:
3921: (c) It is easy to observe that $h(A\cup B)=h(A)\cup h(B)$ for all
3922: $A,B\in L^{\Sigma ^{\ast }}.$ This together with the induction
3923: hypothesis as well as Lemmas 2.5 and 2.6 yields
3924: $$\lceil \gamma(\Delta(\alpha_1 + \alpha_2))\rceil\wedge L(h(\alpha _{1}+\alpha
3925: _{2}))(t) =\lceil \gamma(\Delta(\alpha_1 + \alpha_2))\rceil\wedge
3926: L(h(\alpha _{1})+h(\alpha _{2}))(t)$$
3927: $$=\lceil \gamma(\Delta(\alpha_1 + \alpha_2))\rceil\wedge\lceil \gamma(\Delta(\alpha_1 + \alpha_2))\rceil\wedge
3928: [L(h(\alpha _{1}))(t)\vee L(h(\alpha _{2}))(t)]$$ $$\leq [\lceil
3929: \gamma(\Delta(\alpha_1 + \alpha_2))\rceil\wedge
3930: L(h(\alpha_1))(t)]\wedge [\lceil \gamma(\Delta(\alpha_1 +
3931: \alpha_2))\rceil\wedge L(h(\alpha_2))(t)]$$ $$\leq [\lceil
3932: \gamma(\Delta(\alpha_1))\rceil\wedge L(h(\alpha_1))(t)]\wedge
3933: [\lceil \gamma(\Delta(\alpha_2))\rceil\wedge L(h(\alpha_2))(t)]$$
3934: $$h(L(\alpha _{1}))(t)\vee h(L(\alpha _{2}))(t)$$
3935: $$=h(L(\alpha _{1})\cup L(\alpha _{2}))(t)$$
3936: $$=h(L(\alpha _{1}+\alpha _{2}))(t).$$
3937:
3938: (d) For any $t\in \Gamma ^{\ast },$ Lemmas 2.5, 2.6 and 7.3 enable
3939: us to assert that $$\lceil \gamma(\Delta(\alpha_1\cdot
3940: \alpha_2))\rceil\wedge L(h(\alpha_1\cdot \alpha_2))(t)=\lceil
3941: \gamma(\Delta(\alpha_1\cdot \alpha_2))\rceil\wedge
3942: L(h(\alpha_1)\cdot h(\alpha_2))(t)$$ $$=\lceil
3943: \gamma(\Delta(\alpha_1\cdot \alpha_2))\rceil\wedge
3944: L(h(\alpha_1))L(h(\alpha_2))(t)$$
3945: $$=\lceil \gamma(\Delta(\alpha_1\cdot
3946: \alpha_2))\rceil\wedge \vee\{L(h(\alpha_1))(t_1)\wedge
3947: L(h(\alpha_2))(t_2):t_1t_2=t\}$$
3948: $$\leq \vee\{\lceil \gamma(\Delta(\alpha_1\cdot
3949: \alpha_2))\rceil\wedge L(h(\alpha_1))(t_1)\wedge
3950: L(h(\alpha_2))(t_2):t_1t_2=t\}$$
3951: $$=\vee\{\lceil \gamma(\Delta(\alpha_1\cdot
3952: \alpha_2))\rceil\wedge (\lceil
3953: \gamma(\Delta(\alpha_1))\rceil\wedge L(h(\alpha_1))(t_1))\wedge
3954: (\lceil \gamma(\Delta(\alpha_2))\rceil\wedge
3955: L(h(\alpha_2))(t_2)):t_1t_2=t\}$$
3956: $$\leq \vee\{\lceil \gamma(\Delta(\alpha_1\cdot
3957: \alpha_2))\rceil\wedge h(L(\alpha_1))(t_1)\wedge
3958: h(L(\alpha_2))(t_2):t_1t_2=t\}.$$ Furthermore, by using Lemmas
3959: 2.5, 2.6 and 7.3 again we obtain $$\lceil
3960: \gamma(\Delta(\alpha_1\cdot \alpha_2))\rceil\wedge
3961: h(L(\alpha_1))(t_1)\wedge h(L(\alpha_2))(t_2)=\lceil
3962: \gamma(\Delta(\alpha_1\cdot \alpha_2))\rceil\wedge
3963: (\vee\{L(\alpha_1)(s_1):h(s_1)=t_1\})$$
3964: $$\wedge (\vee\{L(\alpha_2)(s_2):h(s_2)=t_2\})$$
3965: $$\leq \vee\{L(\alpha_1)(s_1)\wedge L(\alpha_2)(s_2):h(s_1)=t_1\ {\rm and}\
3966: h(s_2)=t_2\}.$$ Therefore, it follows that
3967: $$\lceil \gamma(\Delta(\alpha_1\cdot
3968: \alpha_2))\rceil\wedge L(h(\alpha_1\cdot \alpha_2))(t)\leq
3969: \vee\{L(\alpha_1)(s_1)\wedge$$ $$L(\alpha_2)(s_2):h(s_1)=t_1,
3970: h(s_2)=t_2\ {\rm and}\ t_1t_2=t\}$$
3971: $$=\vee\{L(\alpha_1)(s_1)\wedge L(\alpha_2)(s_2):h(s_1s_2)=t\}$$
3972: $$=h(L(\alpha_1)L(\alpha_2))(t)$$ $$=h(L(\alpha_1\alpha_2))(t).$$
3973:
3974:
3975: (e) For every $t\in \Gamma ^{\ast },$ Lemmas 2.5, 2.6 and 7.3
3976: guarantee that $$\lceil \gamma (\Delta (\alpha ^{\ast }))\rceil
3977: \wedge L(h(\alpha ^{\ast }))(t)=\lceil \gamma (\Delta (\alpha
3978: ^{\ast }))\rceil \wedge L((h(\alpha ))^{\ast })(t)$$
3979: $$=\lceil \gamma (\Delta (\alpha ^{\ast }))\rceil \wedge
3980: (L(h(\alpha )))^{\ast }(t)$$
3981: $$=\lceil \gamma (\Delta (\alpha ^{\ast }))\rceil \wedge \vee
3982: \{\wedge _{i=1}^{n}L(h(\alpha ))(t_{i}):n\geq 0,t_{1},...,t_{n}\in
3983: \Gamma ^{\ast },t_{1}...t_{n}=t\}$$
3984: $$\leq \vee \{\lceil \gamma (\Delta (\alpha ^{\ast }))\rceil \wedge
3985: \wedge _{i=1}^{n}L(h(\alpha ))(t_{i}):n\geq 0,t_{1},...,t_{n}\in
3986: \Gamma ^{\ast },t_{1}...t_{n}=t\}$$
3987: $$=\vee \{\lceil \gamma (\Delta (\alpha ))\rceil \wedge \wedge
3988: _{i=1}^{n}(\lceil \gamma (\Delta (\alpha ))\rceil \wedge
3989: L(h(\alpha ))(t_{i})):n\geq 0,t_{1},...,t_{n}\in \Gamma ^{\ast
3990: },t_{1}...t_{n}=t\}$$
3991: $$\leq \vee \{\lceil \gamma (\Delta (\alpha ))\rceil \wedge \wedge
3992: _{i=1}^{n}h(L(\alpha ))(t_{i}):n\geq 0,t_{1},...,t_{n}\in \Gamma
3993: ^{\ast },t_{1}...t_{n}=t\}.$$ On the other hand, we have
3994: $$\lceil \gamma (\Delta (\alpha ))\rceil \wedge \wedge
3995: _{i=1}^{n}h(L(\alpha ))(t_{i})=\lceil \gamma (\Delta (\alpha
3996: ))\rceil \wedge \wedge _{i=1}^{n}(\vee \{L(\alpha
3997: )(s_{i}):h(s_{i})=t_{i}\})$$
3998: $$\leq \vee \{\wedge _{i=1}^{n}L(\alpha )(s_{i}):h(s_{i})=t_{i}\
3999: (i=1,...,n)\}.$$ This further yields
4000: $$\lceil \gamma (\Delta (\alpha ^{\ast }))\rceil \wedge L(h(\alpha
4001: ^{\ast }))(t)\leq \vee \{\wedge _{i=1}^{n}L(\alpha )(s_{i}):n\geq
4002: 0,h(s_{i})=t_{i}\ (i=1,...,n)\ {\rm and}\ t=t_{1}...t_{n}\}$$
4003: $$=\vee \{\wedge _{i=1}^{n}L(\alpha )(s_{i}):n\geq
4004: 0,h(s_{1}...s_{n})=t\}$$
4005: $$=\vee \{L(\alpha )^{\ast }(s):h(s)=t\}$$
4006: $$=h((L(\alpha ))^{\ast })(t)$$
4007: $$=h(L(\alpha ^{\ast }))(t).$$
4008:
4009: What remains is to prove that (ii) implies (i) in (3). This needs
4010: indeed to show that the distributivity of $\wedge$ over $\vee$ is
4011: derivable from the statement (ii). Suppose that $a,b,c\in L.$ We
4012: choose an symbol $\sigma \in \Sigma $ and an symbol $\gamma \in
4013: \Gamma ,$ and define $h(\sigma )=\varepsilon $ and $h(\sigma
4014: ^{\prime })=\gamma $ for every $\sigma ^{\prime }\in \Sigma
4015: -\{\sigma \}.$ We further set $\alpha _{1}=a.\sigma $ and $\alpha
4016: _{2}=b.\varepsilon +c.\sigma .$ Then
4017: $$L(\alpha _{1}.\alpha
4018: _{2})(\sigma )=\left\{
4019: \begin{array}{c}
4020: a\wedge b\ {\rm if }\ n=1, \\
4021: a\wedge c\ {\rm if }\ n=2, \\
4022: 0\ {\rm otherwise,}
4023: \end{array}
4024: \right. $$ and
4025: $$h(L(\alpha _{1}.\alpha _{2}))(\varepsilon )=\vee
4026: _{n=0}^{\infty }L(\alpha _{1}.\alpha _{2})(\sigma ^{n})$$
4027: $$=(a\wedge b)\vee (a\wedge c).$$
4028: On the other hand, we have
4029: $$L(h(\alpha _{1}.\alpha
4030: _{2}))(\varepsilon )=L((a.\varepsilon ).(b.\varepsilon
4031: +c.\varepsilon ))(\varepsilon )$$
4032: $$=L(a.\varepsilon )(\varepsilon
4033: )\wedge L(b.\varepsilon +c.\varepsilon )(\varepsilon )$$
4034: $$=a\wedge
4035: (b\vee c).$$
4036:
4037: From (ii) we know that $h(L(\alpha _{1}.\alpha _{2}))(\varepsilon
4038: )=L(h(\alpha _{1}.\alpha _{2}))(\varepsilon ).$ This indicates
4039: that $ (a\wedge b)\vee (a\wedge c)=a\wedge (b\vee c).\heartsuit$
4040:
4041: \bigskip\
4042:
4043: \textbf{8. Pumping Lemma for Orthomodular Lattice-Valued Regular
4044: Languages}
4045:
4046: \smallskip\
4047:
4048: The pumping lemma in the classical automata theory is a powerful
4049: tool to show that certain languages are not regular, and it
4050: exposes some limitations of finite automata. The purpose of this
4051: section is to establish a generalization of the pumping lemma for
4052: orthomodular lattice-valued languages. It is worth noting that the
4053: following orthomodular lattice-valued version of pumping lemma is
4054: given for the commutative regularity $CReg_\Sigma$. In general,
4055: the pumping lemma is not valid for the predicate $Reg_\Sigma$.
4056:
4057: \smallskip\
4058:
4059: \textbf{Theorem 8.1.} (The pumping lemma) Let $\ell =<L,\leq
4060: ,\wedge ,\vee ,\perp ,0,1>$ be an orthomodular lattice, and let
4061: $\rightarrow =\rightarrow _{3}.$ Then for any $A\in L^{\Sigma
4062: ^{\ast }},$ if $Range(A)$ is finite, then
4063: $$\stackrel{\ell }{\models}CReg_{\Sigma }(A)\rightarrow (\exists n\geq 0)(\forall s\in
4064: \Sigma ^{\ast })[s\in A\wedge |s|\geq n\rightarrow $$
4065: $$(\exists u,v,w\in \Sigma ^{\ast })(s=uvw\wedge |uv|\leq n\wedge
4066: |v|\geq 1\wedge (\forall i\geq 0)(uv^{i}w\in A))],$$ where for any
4067: word $t=\sigma _{1}...\sigma _{k}\in \Sigma ^{\ast },$ $|t|$
4068: stands for the length $n$ of $t.$
4069:
4070: \smallskip\
4071:
4072: \textbf{Proof.} For simplicity, we use $X(s,n)$ to mean the
4073: statement that $u,v,w\in \Sigma ^{\ast },\ s=uvw,\ |uv|\leq n,\
4074: {\rm and}\ |v|\geq 1$ for each $s\in \Sigma^{\ast}$ and $n\geq 0$.
4075: Then it suffices to show that
4076: $$\lceil CReg_{\Sigma }(A)\rceil \leq \vee _{n\geq 0}\wedge _{s\in
4077: \Sigma ^{\ast },|s|\geq n}(A(s)\rightarrow \vee _{X(s,n)}\wedge
4078: _{i\geq 0}A(uv^{i}w)).$$ From Definition 3.3 we know that
4079: $$\lceil CReg_{\Sigma }(A)\rceil =\vee _{\Re \in \mathbf{A}(\Sigma
4080: ,\ell )}(\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil
4081: A\equiv rec_{\Re }\rceil ).$$ Thus, we only need to prove that for
4082: any $\Re \in \mathbf{A}(\Sigma ,\ell ),$
4083: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil A\equiv
4084: rec_{\Re }\rceil \leq \vee _{n\geq 0}\wedge _{s\in \Sigma ^{\ast
4085: },|s|\geq n}(A(s)\rightarrow $$ $$\vee _{X(s,n)}\wedge _{i\geq
4086: 0}A(uv^{i}w)).$$
4087:
4088: Let $Q$ be the set of states of $\Re .$ First, it holds that for
4089: any $s\in \Sigma ^{\ast }$ with $|s|\geq |Q|,$ $$(1)\ \ \ \ \ \ \
4090: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ rec_{\Re
4091: }(s)\leq \vee _{X(s,n)}\wedge _{i\geq 0}rec_{\Re }(uv^{i}w).\ \ \
4092: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$ In fact, suppose
4093: that $s=\sigma _{1}...\sigma _{k}.$ Then
4094: $$(2)\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ rec_{\Re }(s)=\vee _{q_{0},q_{1},...,q_{k}}[I(q_{0})\wedge
4095: T(q_{k})\wedge \wedge _{i=0}^{k-1}\delta (q_{i},\sigma
4096: _{i+1},q_{i+1})].\ \ \ \ \ \ \ \ \ \ \ \ \ \ $$ Therefore, it
4097: amounts to showing that for any $q_{0},q_{1},...,q_{k}\in Q,$
4098: $$ (3)\ \ \ \ \ \ \ \ \ \ \ \ \ I(q_{0})\wedge T(q_{k})\wedge \wedge _{i=0}^{k-1}\delta
4099: (q_{i},\sigma _{i+1},q_{i+1})\leq \vee _{X(s,n)}\wedge _{i\geq
4100: 0}rec_{\Re }(uv^{i}w).\ \ \ \ \ \ \ \ \ \ \ \ $$ Since $k=|s|\geq
4101: |Q|,$ there are two identical states among $
4102: q_{0},q_{1},...,q_{|Q|};$ in other words, there are $m\geq 0$ and
4103: $n>0$ such that $m+n\leq |Q|$ and $q_{m}=q_{m+n}.$ We set
4104: $u_{0}=\sigma _{1}...\sigma _{m},$ $v_{0}=\sigma _{m+1}...\sigma
4105: _{m+n},$ and $w_{0}=\sigma _{m+n+1}...\sigma _{k}.$ Then
4106: $s=u_{0}v_{0}w_{0},$ $|u_{0}v_{0}|=m+n\leq |Q|, $ $|v|=n\geq 1,$
4107: and $$(4) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \vee
4108: _{X(s,n)}\wedge _{i\geq 0}rec_{\Re }(uv^{i}w)\geq \wedge _{i\geq
4109: 0}rec_{\Re }(u_{0}v_{0}^{i}w_{0}).\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4110: \ \ \ \ \ \ \ \ \ \ \ \ $$ From the definition of $rec_{\Re },$ it
4111: is easy to see that for all $i\geq 0, $ $$(5)\ \ \ \ \ \ \ \ \ \ \
4112: \ \ rec_{\Re }(u_{0}v_{0}^{i}w_{0})\geq \lceil path_{\Re
4113: }(q_{0}\sigma _{1}q_{1}...\sigma _{m}q_{m}\ \ \ \ \ \ \ \ \ \ \ \
4114: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4115: \ \ \ \ \ \ \ \ \ $$
4116: $$(\sigma _{m+1}q_{m+1}...\sigma _{m+n}q_{m+n})^{i}\sigma
4117: _{m+n+1}q_{m+n+1}...\sigma _{k}q_{k})\rceil $$
4118: $$=I(q_{0})\wedge T(q_{k})\wedge \wedge
4119: _{j=0}^{m+n-1}\delta (q_{j},\sigma _{j+1},q_{j+1})\wedge \wedge
4120: _{l=1}^{i-1}[\delta (q_{m+n},\sigma _{m+1},q_{m+1})\wedge $$
4121: $$\wedge _{j=m+1}^{m+n-1}\delta (q_{j},\sigma
4122: _{j+1},q_{j+1})]\wedge \wedge _{j=m+n}^{k-1}\delta (q_{j},\sigma
4123: _{j+1},q_{j+1})$$
4124: $$=I(q_{0})\wedge T(q_{k})\wedge \wedge _{j=0}^{k-1}\delta
4125: (q_{j},\sigma _{j+1},q_{j+1})$$ because $q_{m+n}=q_{m}$ and
4126: $\delta (q_{m+n},\sigma _{m+1},q_{m+1})=\delta (q_{m},\sigma
4127: _{m+1},q_{m+1}).$ Thus, by combining (4) and (5), we obtain (3)
4128: which, together with (2), yields (1).
4129:
4130: Now we use Lemmas 2.11(1) and (3) and obtain
4131: $$\vee _{X(s,|Q|)}\wedge _{i\geq 0}rec_{\Re }(uv^{i}w)\rightarrow
4132: \vee _{X(s,|Q|)}\wedge _{i\geq 0}A(uv^{i}w)\geq \lceil \gamma
4133: (atom(\Re )\cup r(A)\rceil \wedge $$ $$\wedge _{X(s,|Q|)}(\wedge
4134: _{i\geq 0}rec_{\Re }(uv^{i}w)\rightarrow \wedge _{i\geq
4135: 0}A(uv^{i}w))$$ $$\geq \lceil \gamma (atom(\Re )\cup r(A)\rceil
4136: \wedge \wedge _{X(s,|Q|)}\wedge _{i\geq 0}(rec_{\Re
4137: }(uv^{i}w)\rightarrow A(uv^{i}w))$$
4138: $$\geq \lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \wedge
4139: _{t\in \Sigma ^{\ast }}(rec_{\Re }(t)\rightarrow A(t))$$
4140: $$=\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil rec_{\Re
4141: }\subseteq A\rceil .$$ Furthermore, from the above inequality we
4142: have
4143: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil rec_{\Re
4144: }\equiv A\rceil =\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge
4145: \lceil A\subseteq rec_{\Re }\rceil \wedge \lceil rec_{\Re
4146: }\subseteq A\rceil $$
4147: $$=\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \wedge _{s\in
4148: \Sigma ^{\ast }}(A(s)\rightarrow rec_{\Re }(s))\wedge \lceil
4149: rec_{\Re }\subseteq A\rceil $$
4150: $$\leq \lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \wedge
4151: _{s\in \Sigma ^{\ast },|s|\geq |Q|}(A(s)\rightarrow rec_{\Re
4152: }(s))\wedge \lceil rec_{\Re }\subseteq A\rceil $$
4153: $$=\wedge _{s\in \Sigma ^{\ast },|s|\geq |Q|}(\lceil \gamma
4154: (atom(\Re )\cup r(A)\rceil \wedge (A(s)\rightarrow rec_{\Re
4155: }(s))\wedge \lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil
4156: rec_{\Re }\subseteq A\rceil )$$
4157: $$\leq \wedge _{s\in \Sigma ^{\ast },|s|\geq |Q|}(\lceil \gamma
4158: (atom(\Re )\cup r(A)\rceil \wedge (A(s)\rightarrow rec_{\Re
4159: }(s))$$
4160: $$\wedge (\vee _{X(s,|Q|)}\wedge _{i\geq 0}rec_{\Re
4161: }(uv^{i}w)\rightarrow \vee _{X(s,|Q|)}\wedge _{i\geq
4162: 0}A(uv^{i}w))).$$ Then from (1) it follows that
4163: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil rec_{\Re
4164: }\equiv A\rceil \leq \wedge _{s\in \Sigma ^{\ast },|s|\geq
4165: |Q|}(\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge$$
4166: $$(A(s)\rightarrow \vee _{X(s,|Q|)}\wedge _{i\geq 0}rec_{\Re
4167: }(uv^{i}w))\wedge$$ $$(\vee _{X(s,|Q|)}\wedge _{i\geq 0}rec_{\Re
4168: }(uv^{i}w)\rightarrow \vee _{X(s,|Q|)}\wedge _{i\geq
4169: 0}A(uv^{i}w))).$$ By using Lemmas 2.11(1) and (3) we know that
4170: $$\lceil \gamma (atom(\Re )\cup r(A)\rceil \wedge \lceil rec_{\Re
4171: }\equiv A\rceil \leq \wedge _{s\in \Sigma ^{\ast },|s|\geq
4172: |Q|}(A(s)\rightarrow \vee _{X(s,|Q|)}\wedge _{i\geq
4173: 0}A(uv^{i}w))$$
4174: $$\leq \vee _{n\geq 0}\wedge _{s\in \Sigma ^{\ast },|s|\geq
4175: n}(A(s)\rightarrow \vee _{X(s,n)}\wedge _{i\geq 0}A(uv^{i}w)),$$
4176: and this completes the proof.$\heartsuit$
4177:
4178: \bigskip\
4179:
4180: \textbf{9. Conclusion}
4181:
4182: \smallskip\
4183:
4184: It is argued that a theory of computation based on quantum logic
4185: has to be established as a logical foundation of quantum
4186: computation. This paper is the first one of a series of papers
4187: toward such a new theory. Quantum logic is treated as an
4188: orthomodular lattice-valued logic in this paper, and the aim of
4189: the paper is to develop elementally a theory of finite automata
4190: based on such a logic by employing the semantical analysis
4191: approach. The notions of orthomodular lattice-valued finite
4192: automaton and regular language are introduced. Some modifications
4193: of orthomodular lattice-valued automaton are presented, including
4194: the orthomodular lattice-valued generalizations of deterministic
4195: and nondeterministic automata and automata with
4196: $\varepsilon-$moves, and their equivalence are thoroughly
4197: analyzed. We also examine the closure properties of orthomodular
4198: lattice-valued regular languages under various operations. The
4199: concept of orthomodular lattice-valued regular expressions is
4200: proposed, and the Kleene theorem concerning the equivalence
4201: between finite automata and regular expressions is generalized
4202: within the framework of quantum logic. Also, an orthomodular
4203: lattice-valued version of the pumping lemma is found. Furthermore,
4204: a theory of pushdown automata or Turing machines based on quantum
4205: logic will be developed in the continuations of the present paper.
4206:
4207: In the development of automata theory based on quantum logic, some
4208: essential differences between the computation theory established
4209: by using the classical Boolean logic as the underlying logical
4210: tool and that whose meta-logic is quantum logic have been
4211: discovered. First, it is found that the proofs of some even very
4212: basic properties of automata appeal an essential application of
4213: the distributivity for the lattice of truth values of the
4214: underlying logic. This indicates that these properties holds only
4215: in Boolean logic but not in quantum logic. We believe that there
4216: are also many fundamental properties of pushdown automata and
4217: Turing machines whose universal validity requires the
4218: distributivity of meta-logic. In a sense, this observation
4219: provides us with a set of negative results in the theory of
4220: computation based on quantum logic. These negative results might
4221: hints some limitations of quantum computers. More explicitly, some
4222: methods based on certain properties of classical automata maybe
4223: have been successfully used in the implementation of classical
4224: computer systems, but they do not apply to quantum computers, or
4225: at least they are only conditionally effective for quantum
4226: computers. On the other hand, although these negative results are
4227: found in the computation theory based on quantum logic, it seems
4228: that some similar negative results exist in other mathematical
4229: theories based on nonclassical logics. This stimulates us to
4230: consider the problem of a logical revisit to mathematics. Various
4231: classical mathematical results have been established based upon
4232: classical logic, and sometimes, their universal validity can only
4233: be established by exploiting the full power of classical logic.
4234: Mathematicians usually use logic implicitly in their reasoning,
4235: and they do not seriously care which logical laws they have
4236: employed. But from a logician's point of view, it is very
4237: interesting to determine how strong a logic we need to validate a
4238: given mathematical theorem, and which logic guarantees this
4239: theorem and which does not among the large population of
4240: nonclassical logics. To be more explicit and also for a
4241: comparison, let us present a short excerpt from A. Heyting [He63,
4242: page 3]:
4243: $$$$
4244: \textit{"It may happen that for the proof of a theorem we do not
4245: need all the axioms, but only some of them. Such a theorem is true
4246: not only for models of the whole system, but also for those of the
4247: smaller system which contains only the axioms used in the proof.
4248: Thus it is important in an axiomatic theory to prove every theorem
4249: from the least possible set of axioms."}
4250: $$$$
4251: We now are in a similar situation. The difference between our case
4252: and A. Heyting's one is that we are concerned with the limitation
4253: or redundance of power of the logic underlying an axiomatic
4254: theory, whereas he considered that of axioms themselves. It seems
4255: that the semantical analysis approach provides a nice framework
4256: for solving this problem, much more suitable than a
4257: proof-theoretical approach.
4258:
4259: As stated above, some fundamental properties of automata are not
4260: universally valid in quantum logic due to lack of distributivity.
4261: However, a certain commutativity are able to regain a local
4262: distributivity, and to give further a partial validity of these
4263: properties in the theory of automata based on quantum logic. One
4264: typical example of such properties is the equivalence of automata
4265: and their various modifications. It is well-known that one
4266: important witness for the Church-Turing thesis which asserts the
4267: Turing machine is a general model of computation is that various
4268: extensions of the Turing machine are all equivalent to itself. The
4269: fact that the equivalence between automata and their modifications
4270: depends upon the commutativity of their basic actions suggests us
4271: to guess that the equivalence between the Turing machine and some
4272: of its extensions may also need a support from a certain
4273: commutativity. In the introduction, we already gave a physical
4274: interpretation to the role of commutativity based on the
4275: Heisenberg uncertainty principle, and pointed out that an
4276: interesting connection may reside between the Heisenberg
4277: uncertainty principle and the Church-Turing thesis. If this is
4278: true, then it will give once again an evidence to the unity of the
4279: whole science and to the fact that science is not only a simple
4280: union of various subjects.
4281:
4282: \bigskip\
4283:
4284: \textbf{References}
4285:
4286: \smallskip\
4287:
4288: [Ba85] J. Barwise, in: J. Barwise and S. Feferman, eds.,
4289: \textit{Model Theoretic Logics}, Springer-Verlag, Berlin, 1985,
4290: pp. ?.
4291:
4292: [Be80] P. A. Benioff, The computer as a physical system: a
4293: microscopic quantum mechanical Hamiltonian model of computer s as
4294: represented by Turing machines, \textit{Journal of Statistical
4295: Physics} \textbf{22}(1980)563-591.
4296:
4297: [Ben73] C. H. Bennet, Logical reversibility of computation,
4298: \textit{IBM Journal of Research and Development}
4299: \textbf{17}(1973)525-532.
4300:
4301: [BN36] G. Birkhoff and J. von Neumann, The logic of quantum
4302: mechanics, \textit{Annals of Mathematics},
4303: \textbf{37}(1936)823-843.
4304:
4305: [BH00] G. Bruns and J. Harding, Algebraic aspects of orthomodular
4306: lattices, in: B. Coecke, D. Moore and A. Wilce (eds.),
4307: \textit{Current Research in Operational Quantum Logic: Algebras,
4308: Categories, Languages,} Kluwer, Dordrecht, 2000, pp. 37-65.
4309:
4310: [CZ95] J. I. Cirac and P. Zoller, Quantum computation with cold
4311: trapped ions, \textit{Physical Review Letters}
4312: \textbf{74}(1995)4091-4094.
4313:
4314: [CM00] J. P. Crutchfield and C. Moore, Quantum automata and
4315: quantum grammar, \textit{Theoretical Computer Science}
4316: \textbf{237}(2000)275-306.
4317:
4318: [DC81] M. L. Dalla Chiara, Some meta-logical pathologies of
4319: quantum logic, in: E. Beltrametti and B. C. van Fraassen (eds.),
4320: \textit{Current Issues in Quantum Logics}, Plenum, New York, 1981,
4321: pp. 147-159.
4322:
4323: [DC86] M. L. Dalla Chiara, Quantum logic, in: D. Gabbay and E.
4324: Guenthner (eds.), \textit{Handbook of Philosophical Logic, volume
4325: III: Alternatives to Classical Logic}, D. Reidel Publishing
4326: Company, Dordrecht, 1986, pp. 427-469. Also see arXiv:
4327: quant-ph/0101028 for an extended and revised version.
4328:
4329: [De85] D. Deutsch, Quantum theory, the Church-Turing principle and
4330: the universal quantum computer, \textit{Proc. Roy. Soc. Lond}.
4331: \textbf{A400}(1985)97-117.
4332:
4333: [De89] D. Deutsch, Quantum computational networks, \textit{Proc.
4334: Roy. Soc. Lond}. \textbf{A425}(1989)73-90.
4335:
4336: [Di78] J. Dieudonne, The current trend of pure mathematics,
4337: \textit{Advances in Mathematics} \textbf{27}(1978)235-255.
4338:
4339: [E74] S. Eilenberg, \textit{Automata, Languages, and Machines,
4340: volume A}, Academic Press, New York, 1974.
4341:
4342: [Fe82] R. P. Feynman, Simulating physics with computers, \textit{
4343: International Journal of Theoretical Physics},
4344: \textbf{21}(1982)467-488.
4345:
4346: [Fe86] R. P. Feynman, Quantum mechanical computers,
4347: \textit{Foundations of Physics,} \textbf{16}(1986)507-531.
4348:
4349: [Fi70] P. D. Finch, Quantum logic as an implication algebra,
4350: \textit{Bulletin of Australian Mathematical Society}
4351: \textbf{2}(1970)101-106.
4352:
4353: [Gr96] L. K. Grover, A fast quantum mechanical algorithm for
4354: database search, in: \textit{Proceedings of the 28th ACM STOC,}
4355: 1996, pp. 212-219.
4356:
4357: [Gu00] S. Gudder, Basic properties of quantum automata,
4358: \textit{Foundations of Physics}, \textbf{30}(2000)301-319.
4359:
4360: [Ha82] W. Hatcher, \textit{The Logical Foundations of
4361: Mathematics}, Pergamon, Oxford (1982).
4362:
4363: [He63] A. Heyting, \textit{Axiomatic Projective Geometry,}
4364: North-Holland, Amsterdam, 1963.
4365:
4366: [Ho89] H. Hodes, Three-valued logics - an introduction, a
4367: comparison of various logical lexica, and some philosophical
4368: remarks, \textit{Annals of Pure and Applied Logic},
4369: \textbf{43}(1989)99-145.
4370:
4371: [Hu37] K. Husimi, Studies on the foundations of quantum mechanics
4372: I, \textit{Proceedings of the Physicomath. Soc. of Japan}
4373: \textbf{19}(1937)766-789.
4374:
4375: [HMP75] L. Herman, E. Marsden and R. Piziak, Implication
4376: connectives in orthomodular lattices, \textit{Notre Dame J. Formal
4377: Logic} \textbf{16}(1975)305-328.
4378:
4379: [HU79] J. E. Hopcroft and J. D. Ullman, \textit{Introduction to
4380: Automata Theory, Languages, and Computation,} Addison-Wesley,
4381: Reading, 1979.
4382:
4383: [Ka74] G. Kalmbach, Orthomodular logic, \textit{Zeitschr. f. math.
4384: Logik und Grundlagen d. Math} \textbf{20}(1974)395-406.
4385:
4386: [Ka83] G. Kalmbach, \textit{Orthomodular Lattices}, Academic
4387: Press, London, 1983.
4388:
4389: [KW97] A. Kondacs and J. Watrous, On the power of quantum finite
4390: state automata, in: \textit{Proc. of the 38th Annual Symposium on
4391: Foundations of Computer Science}, pp.65-75, 1997.
4392:
4393: [Ko97] D. C. Kozen, \textit{Automata and Computability,}
4394: Springer-Verlag, New York, 1997.
4395:
4396: [L93] S. Lloyd, A potentially realizable quantum computer,
4397: \textit{Science} \textbf{261}(1993)1569-1571.
4398:
4399: [Ma90] J. Malinowski, The deduction theorem for quantum logic -
4400: some negative results, \textit{Journal of Symbolic Logic}
4401: \textbf{55}(1990)615-625
4402:
4403: [Mo65] A. Mostowski, \textit{Thirty Years of Foundational Studies:
4404: Lectures on the Development of Mathematical Logic and the Study of
4405: the Foundations of Mathematics in 1930-1964,} Acta Philosophica
4406: Fennica 17, Helsinki, 1965.
4407:
4408: [N62] J. von Neumann, Quantum logics (strict- and
4409: probability-logics), summarized in: J. von Neumann,
4410: \textit{Collected Works, vol. IV}, Macmillan, New York(1962).
4411:
4412: [RS00] J. P. Rawling and S. A. Selesnick, Orthologic and quantum
4413: logic: models and computational elements, \textit{Journal of the
4414: ACM} \textbf{47}(2000)721-751.
4415:
4416: [RR91] L. Rom\'{a}n and B. Rumbos, Quantum logic revisited,
4417: \textit{ Foundations of Physics}, \textbf{21}(1991)727-734.
4418:
4419: [RZ99] L. Rom\'{a}n and R. E. Zuazua, Quantum implication,
4420: \textit{ International Journal of Theoretical Physics,}
4421: \textbf{38}(1999)793-797.
4422:
4423: [RT52] J. B. Rosser and A. R. Turquette, \textit{Many-Valued
4424: Logics,} North-Holland, Amsterdam, 1952.
4425:
4426: [Sc99] K. -G. Schlesinger, Toward quantum mathematics. I. From
4427: quantum set theory to universal quantum mechanics, \textit{Journal
4428: of Mathematics Physics,} \textbf{40}(1999)1344-1358.
4429:
4430: [Sh94] P. W. Shor, Polynomial-time algorithm for prime
4431: factorization and discrete logarithms on quantum computer, in:
4432: \textit{Proc. 35th Annual Symp. on Foundations of Computer
4433: Science, Santa Fe,} IEEE Computer Society Press, 1994.
4434:
4435: [Sv98] K. Svozil, \textit{Quantum Logic,} Springer-Verlag, Berlin,
4436: 1998.
4437:
4438: [T81] G. Takeuti, Quantum set theory, in: E. Beltrametti and B. C.
4439: van Fraassen (eds.), \textit{Current Issues in Quantum Logics},
4440: Plenum, New York, 1981, pp. 303-322.
4441:
4442: [TD88] A. S. Troelstra and D. van Dalen, \textit{Constructivism in
4443: Mathematics: An Introduction,}, volume I, II, North-Holland,
4444: Amsterdam, 1988.
4445:
4446: [VP] V. Vedral and M. B. Plenio, Basics of quantum computation,
4447: \textit{ Prog. Quant. Electron}. \textbf{22}:1(1998).
4448:
4449: [Ya93] A. C. Yao, Quantum circuit complexity, \textit{Proc. of the
4450: 34th Ann. IEEE Symp. on Foundations of Computer Science}, pp.
4451: 352-361, 1993.
4452:
4453: [Yi91] M. S. Ying, Deduction theorem for many-valued inference,
4454: \textit{ Zeitschr. f. math. Logik und Grundlagen d. Math}.
4455: 37:6(1991).
4456:
4457: [Yi92a] M. S. Ying, The fundamental theorem of ultraproduct in
4458: Pavelka's logic, \textit{Zeitschr. f. math. Logik und Grundlagen
4459: d. Math.} 38:2(1992).
4460:
4461: [Yi92b] M. S. Ying, Compactness, the Lowenheim-Skolem property and
4462: the direct product of lattices of truth values, \textit{Zeitschr.
4463: f. math. Logik und Grundlagen d. Math}. 38:4(1992).
4464:
4465: [Yi91-93] M. S. Ying, A new approach for fuzzy topology (I), (II),
4466: (III), \textit{Fuzzy Sets and Systems,} \textbf{39}(1991)303-321;
4467: \textbf{47} (1992)221-232; \textbf{55}(1993)193-207.
4468:
4469: [Yi93] M. S. Ying, Fuzzifying topology based on complete
4470: residuated lattice-valued logic (I), \textit{Fuzzy Sets and
4471: Systems,} \textbf{56}(1993)337-373.
4472:
4473: [Yi94] M. S. Ying, A logic for approximate reasoning, \textit{J.
4474: Symbolic Logic} \textbf{59}(1994).
4475:
4476: [Yi00] M. S. Ying, Automata theory based on quantum logic (I),
4477: (II), \textit{International Journal of Theoretical Physics}
4478: \textbf{39} (2000), 985-995; 2545-2557.
4479:
4480: \end{document}
4481: