Theoretical Computer Science. Hierarchy and equivalence of multi-letter quantum finite automata

Similar documents
Determining the equivalence for one-way quantum finite automata

arxiv:quant-ph/ v2 21 Apr 2007

Simple models of quantum finite automata

Lower Bounds for Generalized Quantum Finite Automata

QUANTUM FINITE AUTOMATA. Andris Ambainis 1 1 Faculty of Computing, University of Latvia,

On the Power of Quantum Queue Automata in Real-time

Running head: Watson-Crick Quantum Finite Automata. Title: Watson-Crick Quantum Finite Automata. Authors: Kingshuk Chatterjee 1, Kumar Sankar Ray 2

On the complexity of minimizing probabilistic and quantum automata

Running head: 2-tape 1-way Quantum Finite State Automata. Title: Quantum Finite State Automata. Authors: Debayan Ganguly 1, Kumar Sankar Ray 2

Theoretical Computer Science. State complexity of basic operations on suffix-free regular languages

FORMAL LANGUAGES, AUTOMATA AND COMPUTATION

An Optimal Lower Bound for Nonregular Languages

On languages not recognizable by One-way Measure Many Quantum Finite automaton

COMPARATIVE ANALYSIS ON TURING MACHINE AND QUANTUM TURING MACHINE

NQP = co-c = P. Electronic Colloquium on Computational Complexity, Report No. 73 (1998)

arxiv:cs/ v1 [cs.cc] 2 Oct 2001

10 Quantum Complexity Theory I

Varieties Generated by Certain Models of Reversible Finite Automata

On Closure Properties of Quantum Finite Automata

Computability and Complexity

Varieties Generated by Certain Models of Reversible Finite Automata

Automata Theory. Lecture on Discussion Course of CS120. Runzhe SJTU ACM CLASS

Results on Transforming NFA into DFCA

State Complexity of Two Combined Operations: Catenation-Union and Catenation-Intersection

Theory of Computation

Discrete quantum random walks

Quantum Computing Lecture 8. Quantum Automata and Complexity

c 1998 Society for Industrial and Applied Mathematics Vol. 27, No. 4, pp , August

Undecidable Problems and Reducibility

State Complexity of Neighbourhoods and Approximate Pattern Matching

On Rice s theorem. Hans Hüttel. October 2001

3515ICT: Theory of Computation. Regular languages

CS5371 Theory of Computation. Lecture 10: Computability Theory I (Turing Machine)

Limitations of Non-deterministic Finite Automata Imposed by One Letter Input Alphabet

A Uniformization Theorem for Nested Word to Word Transductions

The exact complexity of the infinite Post Correspondence Problem

The Turing machine model of computation

DESCRIPTIONAL COMPLEXITY OF NFA OF DIFFERENT AMBIGUITY

Mergible States in Large NFA

Incompleteness Theorems, Large Cardinals, and Automata ov

Combinatorial Interpretations of a Generalization of the Genocchi Numbers

On Stateless Multicounter Machines

Introduction to Languages and Computation

arxiv: v1 [cs.cc] 3 Feb 2019

Minimizing finite automata is computationally hard

Computation Histories

Parallel Turing Machines on a Two-Dimensional Tape

ECS 120 Lesson 15 Turing Machines, Pt. 1

ACS2: Decidability Decidability

Reversal of Regular Languages and State Complexity

An algebraic characterization of unary two-way transducers

Turing Machines Part III

CS5371 Theory of Computation. Lecture 10: Computability Theory I (Turing Machine)

Opleiding Informatica

CS 154, Lecture 2: Finite Automata, Closure Properties Nondeterminism,

Turing machines Finite automaton no storage Pushdown automaton storage is a stack What if we give the automaton a more flexible storage?

CSE355 SUMMER 2018 LECTURES TURING MACHINES AND (UN)DECIDABILITY

Turing Machines, diagonalization, the halting problem, reducibility

Finite Automata. Mahesh Viswanathan

CSC173 Workshop: 13 Sept. Notes

CS21 Decidability and Tractability

Theoretical Computer Science. Complexity of multi-head finite automata: Origins and directions

arxiv:quant-ph/ v2 14 Aug 2001

Lecture Notes On THEORY OF COMPUTATION MODULE -1 UNIT - 2

arxiv: v2 [quant-ph] 19 Mar 2016

FORMAL LANGUAGES, AUTOMATA AND COMPUTABILITY

Ultrametric Automata with One Head Versus Multihead Nondeterministic Automata

Quantum Finite Automata, Quantum Pushdown Automata & Quantum Turing Machine: A Study

Efficient minimization of deterministic weak ω-automata

Watson-Crick ω-automata. Elena Petre. Turku Centre for Computer Science. TUCS Technical Reports

Nondeterministic State Complexity of Basic Operations for Prefix-Free Regular Languages

Descriptional Complexity of Formal Systems (Draft) Deadline for submissions: April 20, 2009 Final versions: June 18, 2009

Turing Machines A Turing Machine is a 7-tuple, (Q, Σ, Γ, δ, q0, qaccept, qreject), where Q, Σ, Γ are all finite

A shrinking lemma for random forbidding context languages

arxiv:cs/ v1 [cs.cc] 9 Feb 2007

The P versus NP Problem in Quantum Physics

Theory of Computation (IX) Yijia Chen Fudan University

CSCE 551: Chin-Tser Huang. University of South Carolina

Remarks on Separating Words

Notes on State Minimization

Variants of Turing Machine (intro)

By allowing randomization in the verification process, we obtain a class known as MA.

cse303 ELEMENTS OF THE THEORY OF COMPUTATION Professor Anita Wasilewska

Formal Definition of Computation. August 28, 2013

Graph Non-Isomorphism Has a Succinct Quantum Certificate

A Lower Bound of 2 n Conditional Jumps for Boolean Satisfiability on A Random Access Machine

Hierarchy among Automata on Linear Orderings

Midterm Exam 2 CS 341: Foundations of Computer Science II Fall 2018, face-to-face day section Prof. Marvin K. Nakayama

Implementing Competitive Learning in a Quantum System

State Complexity Advantages of Ultrametric Automata

A Lower Bound for Boolean Satisfiability on Turing Machines

On Properties and State Complexity of Deterministic State-Partition Automata

Time-bounded computations

Introduction to Theory of Computing

How to Pop a Deep PDA Matters

P systems based on tag operations

COMP/MATH 300 Topics for Spring 2017 June 5, Review and Regular Languages

The efficiency of identifying timed automata and the power of clocks

CS 208: Automata Theory and Logic

The Power of One-State Turing Machines

Transcription:

Theoretical Computer Science 410 (2009) 3006 3017 Contents lists available at ScienceDirect Theoretical Computer Science journal homepage: www.elsevier.com/locate/tcs Hierarchy and equivalence of multi-letter quantum finite automata Daowen Qiu a,c, Sheng Yu b, a Department of Computer Science, Zhongshan University, Guangzhou 510275, China b Department of Computer Science, The University of Western Ontario, London, Ontario, N6A 5B7, Canada c SQIG Instituto de Telecomunicações, IST, TULisbon, Av. Rovisco Pais 1049-001, Lisbon, Portugal a r t i c l e i n f o a b s t r a c t Keywords: Quantum computing Multi-letter finite automata Quantum finite automata Equivalence Hierarchy Multi-letter quantum finite automata (QFAs) are a new one-way QFA model proposed recently by Belovs, Rosmanis, and Smotrovs [A. Belovs, A. Rosmanis, J. Smotrovs, Multiletter reversible and quantum finite automata, in: Proceedings of the 13th International Conference on Developments in Language Theory, DLT 2007, Harrachov, Czech Republic, in: Lecture Notes in Computer Science, vol. 4588, Springer, Berlin, 2007, pp. 60 71], and they showed that multi-letter QFAs can accept with no error some regular languages ((a+b) b) that are unacceptable by the one-way QFAs. In this paper, we continue to study multiletter QFAs. We mainly focus on two issues: (1) we show that ( + 1)-letter QFAs are computationally more powerful than -letter QFAs, that is, ( + 1)-letter QFAs can accept some regular languages that are unacceptable by any -letter QFA. A comparison with the one-way QFAs is made by some examples; (2) we prove that a 1 -letter QFA A 1 and another 2 -letter QFA A 2 are equivalent, if and only if, they are (n 1 +n 2 ) 4 + 1-equivalent, and the time complexity of determining the equivalence of two multi-letter QFAs using this method is O(n 12 + 2 n 4 +n 8 ), where n 1 and n 2 are the numbers of states of A 1 and A 2, respectively, and = max( 1, 2 ). Some other issues are addressed for further consideration. 2009 Elsevier B.V. All rights reserved. 1. Introduction Quantum computing is an intriguing and promising research field, which touches on computer science, quantum physics, and mathematics [16,17,11,10]. To a certain extent, quantum computing was motivated by the exponential speed-up of Shor s quantum algorithm for factoring integers in polynomial time [32] and Grover s algorithm of searching in database of size n with only O( n) accesses [15]. Quantum computers the physical devices complying with the rules of quantum mechanics were first considered by Benioff [8], and then suggested by Feynman [14]. By elaborating and formalizing Benioff and Feynman s idea, in 1985, Deutsch [12] re-examined the Church Turing Principle and defined quantum Turing machines (QTMs). Subsequently, Deutsch [13] considered quantum networ models. In 1993, Yao [35] demonstrated the equivalence between QTMs and quantum circuits. Quantum computation from the viewpoint of complexity theory was first studied systematically by Bernstein and Vazirani [7]. Another ind of simpler model of quantum computation is quantum finite automata (QFAs), which can be thought of as theoretical models of quantum computers with finite memory. This ind of computing machines were first studied by Moore This research is supported by the National Natural Science Foundation (Nos. 60573006, 60873055), the Research Foundation for the Doctorial Program of Higher School of Ministry of Education (No. 20050558015), Program for New Century Excellent Talents in University (NCET) of China, and the Natural Science and Engineering Research Council of Canada Grand #OGP0041630. Corresponding author. Tel.: +1 519 661 3715. E-mail addresses: issqdw@mail.sysu.edu.cn (D. Qiu), syu@csd.uwo.ca (S. Yu). 0304-3975/$ see front matter 2009 Elsevier B.V. All rights reserved. doi:10.1016/j.tcs.2009.03.040

D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 3007 and Crutchfield [26], as well as by Kondacs and Watrous [23] independently. Then it was dealt with in depth by Ambainis and Freivalds [1], Brodsy and Pippenger [5], and the other authors (for example, see the references in [16,30]). The study of QFAs is mainly divided in two ways: one is one-way quantum finite automata (1QFAs) whose tape heads only move one cell to the right at each computation step (1QFAs have been extensively studied [4]), and the other is two-way quantum finite automata (2QFAs), in which the tape heads are allowed to move towards the right or left, or to be stationary [23]. (Notably, Amano and Iwama [2] dealt with a decidability problem concerning an intermediate form called 1.5QFAs, whose tape heads are allowed to move right or to be stationary; Hirvensalo [18] investigated a decidability problem related to one-way QFAs.) Furthermore, by considering the number of times of the measurement in a computation, 1QFAs have two different forms: measure-once 1QFAs (MO-1QFAs) proposed by Moore and Crutchfield [26], and, measure-many 1QFAs (MM-1QFAs) studied first by Kondacs and Watrous [23]. MM-1QFAs are strictly more powerful than MO-1QFAs [1,4] (Indeed, a b can be accepted by MM-1QFAs with bounded error but not by any MO-1QFA with bounded error). Due to the unitarity of quantum physics and finite memory of finite automata, both MO-1QFAs and MM-1QFAs can only accept proper subclasses of regular languages with bounded error (e.g., [23,1,5,4]). Indeed, it was shown that the regular language (a + b) b cannot be accepted by any MM-1QFA with bounded error [23]. Recently, Belovs, Rosmanis, and Smotrovs [6] proposed a new one-way QFA model, namely, multi-letter QFAs, that can be thought of as a quantum counterpart of more restricted classical one-way multi-head finite automata (see, for example, [19]). Roughly speaing, a -letter QFA is not limited to seeing only one, the just-incoming input letter, but can see several earlier received letters as well. That is, the quantum state transition which the automaton performs at each step depends on the last letters received. For the other computing principle, it is similar to the usual MO-1QFAs as described above. Indeed, when = 1, it reduces to an MO-1QFA. Any given -letter QFA can be simulated by some +1-letter QFA. However, we will prove that the contrary does not hold. Belovs et al. [6] have already showed that (a + b) b can be accepted by a 2-letter QFA but, as proved in [23], it cannot be accepted by any MM-1QFA with bounded error. By L(QFA ) we denote the class of languages accepted with bounded error by -letter QFAs. In this paper, we will prove that L(QFA ) L(QFA +1 ) for = 1, 2,..., where the inclusion is proper. Therefore, ( + 1)-letter QFAs are computationally more powerful than -letter QFAs. As we now, determining the equivalence for computing models is a very important issue in the theory of classical computation (see, e.g., [27,33,31,9,21,20]). Concerning the problem of determining the equivalence for QFAs, there exists some wor [5] that deals with the simplest case MO-1QFAs. For quantum sequential machines (QSMs), Qiu [28] gave a negative outcome for determining the equivalence of QSMs, and then Li and Qiu [24] further gave a method for determining whether or not any two given QSMs are equivalent. This method applies to determining the equivalence between any two MO-1QFAs and also is different from the previous ones. For the equivalence problem of MM-1QFAs, inspired by the wor of [34,4], Li and Qiu [25] presented a polynomial-time algorithm for determining whether or not any two given MM-1QFAs are equivalent. In this paper, we will give a polynomial-time algorithm for determining whether or not any two given 1 -letter QFA A 1 and 2 -letter QFA A 2 for accepting unary languages are equivalent. More specifically, we prove that two multi-letter QFAs A 1 and A 2, are equivalent if and only if they are (n 1 + n 2 ) 4 + 1-equivalent, where n 1 and n 2 are the numbers of states of A 1 and A 2, respectively, = max( 1, 2 ), and two multi-letter QFAs over the same input alphabet Σ are n-equivalent if and only if the accepting probabilities of A 1 and A 2 are equal for the input strings of length not more than n. This method, generalized appropriately, may apply to dealing with more general cases. The remainder of the paper is organized as follows: In Section 2, we recall the definition of multi-letter QFAs and other related definitions, and some related results are reviewed. In Section 3, we prove that L(QFA ) L(QFA +1 ) for = 1, 2,..., where the inclusion is proper. More precisely, we show that, for 2, regular language (a 1 + a 2 + + a ) a 1 a 2 a 1 cannot be accepted with bounded error by ( 1)-letter QFAs but can be exactly accepted by some -letter QFAs. In addition, we present a number of examples to show the relation between multi-letter QFAs and the usual one-way QFAs. In Section 4, we concentrate on the equivalence issue. After proving some useful lemmas, we prove that a 1 -letter QFA A 1 and another 2 -letter QFA A 2 for accepting unary languages are equivalent if and only if they are (n 1 +n 2 ) 4 + 1-equivalent, and the time complexity of determining the equivalence of two multi-letter DFAs using this method is O(n 12 + 2 n 4 + n 8 ), where n = n 1 + n 2, n 1 and n 2 are the numbers of states of A 1 and A 2, respectively, and = max( 1, 2 ). Finally, in Section 5 we address some related issues for further consideration. In general, symbols will be explained when they first appear. 2. Preliminaries In this section, we briefly review some definitions and related properties that will be used in the sequel. For the details, we refer to [6]. First we recall -letter deterministic finite automata (-letter DFAs). Definition 1 ([6]). A -letter deterministic finite automaton (-letter DFA) is defined by a quintuple (Q, Q acc, q 0, Σ, γ ), where Q is a finite set of states, Q acc Q is the set of accepting states, q 0 Q is the initial state, Σ is a finite input alphabet,

3008 D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 and γ is a transition function that maps Q T to Q, where T = {Λ} Σ and letter Λ / Σ denotes the blan symbol (lie a blan symbol in Turing machines [33]), and T T consists of all strings of length. We describe the computing process of a -letter DFA on an input string x in Σ, where x = σ 1 σ 2 σ n, and Σ denotes the set of all strings over Σ. The -letter DFA has a tape which contains the letter Λ in its first 1 position followed by the input string x. The automaton starts in the initial state q 0 and has reading heads which initially are on the first positions of the tape (clearly, the th head reads σ 1 and the other heads read Λ). Then the automaton transfers to a new state as current state and all heads move right a position in parallel. Now the ( 1)th and th heads point to σ 1 and σ 2, respectively, and the others, if any, to Λ. Subsequently, the automaton transfers to a new state and all heads move to the right. This process does not stop until the th head has read the last letter σ n. The input string x is accepted if and only if the automaton enters an accepting state after its th head reading the last letter σ n. Clearly, -letter DFAs are not more powerful than DFAs. The family of languages accepted by -letter DFAs, for 1, is exactly the family of regular languages. For the sae of readability, we briefly recall the definitions of MO-1QFAs and MM-1QFAs in the following. An MO-1QFA is defined as a quintuple A = (Q, Q acc, ψ 0, Σ, {U(σ )} σ Σ ), where Q is a set of finite states, Q acc Q is the set of accepting states, ψ 0 is the initial state that is a superposition of the states in Q, Σ is a finite input alphabet, and U(σ ) is a unitary matrix for each σ Σ. As usual, we identify Q with an orthonormal base of a complex Euclidean space and every state q Q is identified with a basis vector, denoted by Dirac symbol q (a column vector), and q is the conjugate transpose of q. We describe the computing process for any given input string x = σ 1 σ 2 σ m Σ. At the beginning the machine A is in the initial state ψ 0, and upon reading σ 1. The transformation U(σ 1 ) acts on ψ 0. After that, U(σ 1 ) ψ 0 becomes the current state and the machine reads σ 2. The process continues until the machine has read σ m ending in the state ψ x = U(σ m )U(σ m 1 ) U(σ 1 ) ψ 0. Finally, a measurement is performed on ψ x and the accepting probability p a (x) is equal to p a (x) = ψ x P a ψ x = P a ψ x 2 where P a = q Q acc q q is the projection onto the subspace spanned by { q : q i Q acc }. An MM-1QFA is defined as a 6-tuple A = (Q, Q acc, Q rej, ψ 0, Σ, {U(σ )} σ Σ {$} ), where Q, Q acc Q, ψ 0, Σ, {U(σ )} σ Σ {$} are the same as those in an MO-1QFA defined above,q rej Q represents the set of rejecting states, and $ Σ is a tape symbol denoting the right end-mar. For any input string x = σ 1 σ 2 σ m Σ, the computing process is similar to that of MO-1QFAs except that after every transition, A measures its state with respect to the three subspaces that are spanned by the three subsets Q acc, Q rej, and Q non, respectively, where Q non = Q \ (Q acc Q rej ). In other words, the projection measurement consists of {P a, P r, P n } where P a = q Q acc q q, P r = q Q rej q q, P n = q Q \(Q acc Q rej ) q q. The machine stops after the right end-mar $ has been read. Of course, the machine may also stop before reading $ if the current state of the machine reading some σ i (1 i m) does not contain the states of Q non. Since the measurement is performed after each transition with the states of Q non being preserved, the accepting probability p a (x) and the rejecting probability p r (x) are given as follows (for convenience, we denote $ = σ m+1 ): 2 m+1 1 p a (x) = P au(σ ) (P n U(σ i )) ψ 0, =1 i=1 2 m+1 1 p r (x) = P ru(σ ) (P n U(σ i )) ψ 0. =1 i=1 We further recall the definitions of a group finite automaton (GFA) [5] and a one-way reversible finite automaton (1RFA) [1]. A GFA is a DFA whose state transition function, say δ, satisfies that for any input symbol σ, δ(, σ ) is a one-to-one map on the state set, i.e., a permutation. A 1RFA is defined as an MO-1QFA but restricting the values of its state transition function onto {0, 1}. More specifically, a 1RFA is a DFA whose set of states, input alphabet, and state transition function are Q, Σ, δ, respectively, where δ satisfies that, for any q Q and any σ Σ, there is at most one p Q such that δ(p, σ ) = q. Qiu [29] proved that GFAs and 1RFAs are equivalent, i.e., any GFA can be simulated by a 1RFA and vice-versa. Definition 2 ([6]). A -letter DFA (Q, Q acc, q 0, Σ, γ ) is called a -letter group finite automaton (-letter GFA) if and only if for any string x T the function γ x (q) = γ (q, x) is a bijection from Q to Q. Remar 1. When = 1, a 1-letter DFA is exactly a DFA [31,33,36], and a 1-letter GFA is also the usual GFA [5]. By L(GFA ) and L(DFA ) we denote the classes of all languages accepted by -letter GFAs and by -letter DFAs, respectively. In addition, we denote L(GFA ) = =1 L(GFA ) and L(DFA ) = =1 L(DFA ). In [6] it was shown that L(GFA) L(GFA ) L(DFA) = L(DFA ), (1) where is a proper inclusion. Now we further recall the definition of multi-letter QFAs [6].

D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 3009 Definition 3 ([6]). A -letter QFA A is defined as a quintuple A = (Q, Q acc, ψ 0, Σ, µ) where Q is a set of states, Q acc Q is the set of accepting states, ψ 0 is the initial unit state that is a superposition of the states in Q, Σ is a finite input alphabet, and µ is a function that assigns a unitary transition matrix U w on C Q for each string w ({Λ} Σ), where Q is the cardinality of Q. The computation of a -letter QFA A wors in the same way as the computation of an MO-1QFA, except that it applies unitary transformations corresponding not only to the last letter but the last letters received (lie a -letter DFA). When = 1, it is exactly an MO-1QFA as pointed out before. According to [6], all languages accepted by -letter QFAs with bounded error are regular languages for any. Now we give the probability P A (x) for -letter QFA A = (Q, Q acc, ψ 0, Σ, µ) accepting any input string x = σ 1 σ 2 σ m. From the definition we now that, for any w ({Λ} Σ), µ(w) is a unitary matrix. In terms of the definition of µ, we can define the unitary transition for each string x = σ 1 σ 2 σ m Σ. By µ we mean a map from Σ to the set of all Q -order unitary matrices. Indeed, µ is induced by µ in the following way. For x = σ 1 σ 2 σ m Σ, { µ(λ 1 σ µ(x) = 1 )µ(λ 2 σ 1 σ 2 ) µ(λ m x), if m <, µ(λ 1 σ 1 )µ(λ 2 (2) σ 1 σ 2 ) µ(σ m +1 σ m +2 σ m ), if m, which implies the computing process of A for input string x. As before, we identify the states in Q with an orthonormal basis of the complex Euclidean space C Q, and let P acc denote the projector on the subspace spanned by Q acc. Then we define that P A (x) = ψ 0 µ(x)p acc 2. Definition 4 ([6]). For 1, a DFA contains a C -construction if and only if there are states q 1, q 2, q 3, q 4, q 5 and a string w = σ 1 σ 2 σ of length such that q 2 q 5, and transformation function γ satisfies γ (q 2, σ ) = γ (q 5, σ ) = q 3, γ (q 1, σ 1 σ 1 ) = q 2 and γ (q 4, σ 1 σ 1 ) = q 5. In the above C -construction, if there exists an m > 0 such that γ (q 3, w m 1 ) = q 4, then we call it a D -construction. Proposition 1 ([6]). If there exists a C -construction in a DFA, then there also exists a D -construction in this DFA. Theorem 2 ([6]). The following statements are equivalent: A language L is in L(QFA ), i.e., L is accepted by a -letter QFA with bounded error. The minimal DFA of L contains no C -construction. L is accepted by a -letter GFA. From Theorem 2we now that a language is accepted by a -letter GFA if and only if it is accepted by a -letter QFA with bounded error. For = 1, it was proved by Brodsy and Pippenger [5]. (3) 3. Hierarchy of multi-letter QFAs and some relations In this section, we deal with two issues. In Section 3.1, we consider the hierarchy of multi-letter QFAs and prove that j-letter QFA are strictly more powerful than i-letter QFAs for 1 i < j. In Section 3.2, we attempt to clarify the relations between the families of languages accepted by multi-letter QFAs and MO-1QFAs and also between those by multi-letter QFAs and MM-QFAs. 3.1. Hierarchy of multi-letter QFAs Are -letter QFAs more powerful than ( 1)-letter QFAs for = 1, 2,...? The answer is positive for = 2 as proved in [6]. In this subsection, we demonstrate that -letter QFAs are more powerful than ( 1)-letter QFAs for any 3. Theorem 3. For any 3, there exists a language that can be accepted by a -letter GFA but cannot be accepted by any ( 1)- letter GFA. Proof. We consider the regular language (a 1 +a 2 + +a ) a 1 a 2 a 1 denoted by L over alphabet Σ = {a 1, a 2,..., a }, and we will prove that L satisfies the theorem. First we construct a minimal DFA for L as A = (Q, Σ, q 0, δ, F) where: Q = {q 0, q 1,..., q 1 }; Σ = {a 1, a 2,..., a }; F = {q 1 }; δ is defined as follows: δ(q 0, a 1 ) = q 1 ; δ(q 0, a i ) = q 0 for i = 2, 3,..., ; δ(q 1, a 1 ) = q 1 ; δ(q 1, a 2 ) = q 2 ; δ(q 1, a i ) = q 0 for i = 3, 4,..., ; δ(q l, a l+1 ) = q l+1 and δ(q l, a t ) = q 0 for l = 2, 3,..., 1 and t {2,..., l, l + 2, l + 3,..., }, where we denote q = q 0.

3010 D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 Fig. 1. A state transition diagram of DFA A. Fig. 2. A C 1 -construction in DFA A. Fig. 3. A supposed C -construction. δ(q i, a 1 ) = q 1 for i = 2, 3,..., 1. Fig. 1 depicts the DFA A above described. We prove that A is a minimal DFA. It suffices to prove that, for all states q 0, q 1,..., q 1, any two different states are distinguishable [22]. In other words, for any 0 i, j 1 with i j, there exists w Σ such that exactly one of δ (q i, w) and δ (q j, w) is the accepting state q 1. Indeed, we can divide it into three cases. 1. q i and q 1 for 0 i 2. Tae w = ɛ, empty string. Then δ (q i, ɛ) = q i and δ (q 1, ɛ) = q 1. 2. q 0 and q l for 1 l 2. Tae w = a l+1 a l+2 a 1. Then δ (q 0, w) = q 0 and δ (q l, w) = q 1. 3. q i and q j for 1 i < j 2. Tae w = a j+1 a j+2 a 1. Then δ (q j, w) = q 1 and δ (q i, w) = q 0. Therefore, we have proved that any two different states of q 0, q 1,..., q 1 are distinguishable. Consequently, A is minimal. In fact, we can see that the number of states is minimal from the number of equivalence classes over Σ [22]. This equivalence relation is defined as: for any w 1, w 2 Σ, w 1 w 2 iff for any z Σ, either both w 1 z and w 2 z in L, or neither w 1 z nor w 2 z in L. Then we can divide Σ into the following equivalence classes: [ɛ], [a 1 ], [a 1 a 2 ],, [a 1 a 2 a 1 ]. As a result, is the number of states of the minimal DFA accepting L. In the state transition figure of A, we find a C 1 -construction. In fact, set w = a 2 a 3... a. Since δ(q 0, a i ) = q 0 for i = 2, 3,...,, we get δ (q 0, a 2 a 3 a 1 ) = q 0 and δ(q 0, a ) = q 0. Moreover, δ(q i, a i+1 ) = q i+1 for i = 1, 2,...,, where we denote q = q 0. This C 1 -construction is better described by Fig. 2. By Theorem 2 we conclude that A cannot be accepted by any ( 1)-letter QFA with bounded error. However, we will verify that, in the minima DFA A, there is no C -construction. Therefore, according to Theorem 2 L can be accepted by a -letter QFA with bounded error. Now we chec that there is no C -construction in A. We prove it by contradiction. Indeed, suppose that there is a C - construction depicted by Fig. 3. We divide the proof into the following three cases. 1. q = q 0. σ = a 1 : It is impossible since q 0 cannot be accessed by inputting a 1. σ = a 2 : In this case, q i, q j {q 0, q 2, q 3,..., q 1 }. If one of q i, q j, say q i is q 0, and the other one q j belongs to {q 2, q 3,..., q 1 }, then σ 1 = a j where j 2. Thus, q j 1 = q j 1 and q i 1 = q 0. In succession, we find that q it = q 0, q jt = q 1 for some 2 t. However, there is no σ Σ leading to q 0 and q 1 simultaneously. Therefore, it is impossible. If q i, q j {q 2, q 3,..., q 1 }, then the above case shows that this is impossible either. Consequently, this case does not exist.

D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 3011 Fig. 4. A relation diagram where q = p 2 = r. σ = a s for 3 s : These cases can be similarly verified as above, and we leave the details out here. 2. q = q 1 σ = a 1 : In this case, q i, q j {q 0, q 1, q 2, q 3,..., q 1 }. Similar to the above proof. σ = a s for 2 s : It is clearly impossible. 3. q = q s for 2 s : For any 2 s, there is no σ Σ and two different states p 1 p 2 such that δ(p 1, σ ) = δ(p 2, σ ) = q s. Consequently, there does not exist such a C -construction. Hence, there does not exist a C -construction in A, and therefore, by Theorem 2 L can be accepted by a -letter GFA. From Theorems 2 and 3 we have the following corollary. Corollary 4. For 2, L(QFA 1 ) L(QFA ), where the inclusion is proper. 3.2. Comparison of multi-letter QFAs with others In this subsection, we try to compare the relations between the families of languages accepted by multi-letter QFAs and MO-1QFAs and also between those by multi-letter QFAs and MM-QFAs. First we recall the definition of forbidden construction in a DFA [1]. In a DFA, a forbidden construction means that there exist string x and states p 1 and p 2, p 1 p 2, such that δ (p 1, x) = p 2 and δ (p 2, x) = p 2, where p 2 is neither all-accepting state, nor all-rejecting. A state p is neither all-accepting state, nor all-rejecting whenever there exist w 1, w 2 Σ such that exactly one of δ (p, w 1 ) and δ (p, w 2 ) is an accepting state. Remar 2. Ambainis and Freivalds [1] presented a forbidden construction and showed that, if the minimal DFA for accepting a regular language does not contain a forbidden construction, then this language can be accepted by a one-way reversible finite automaton. In [29], Qiu proved that one-way reversible finite automata are also GFAs and vice versa. Also, Ambainis and Freivalds [1] proved that a regular language is accepted by an MM-1QFA with bounded error and with probability over 7 9 if and only if this language is accepted by a 1RFA and thus by a GFA as well. Next we verify that a forbidden construction implies a C 1 -construction. Proposition 5. In a DFA, if there exists a forbidden construction, then there also exists a C 1 -construction. Proof. Let A = (Q, Q acc, q 0, Σ, δ) be a DFA. Suppose that there is a forbidden construction, that is, there are states p 1, p 2 and x Σ satisfying δ(p 1, x) = p 2 and δ(p 2, x) = p 2. Suppose that x = σ 1 σ 2... σ. Then there are states q 1, q 2,..., q and r 1, r 2,..., r with q = p 2 = r such that δ(p 1, σ 1 ) = q 1, δ(q, σ 1 ) = r 1, δ(q i, σ i+1 ) = q i+1, where i = 1, 2,..., 1. This relation can be described by Fig. 4. Since p 1 p 2 = q but q = r, there exists q i = r i but q i 1 r i 1. Therefore, we have δ(q i 1, σ i ) = q i = r i and δ(r i 1, σ i ) = r i = r i, which is a C 1 -construction. By Remar 2 and Theorem 2 we obtain the following corollary. Proposition 6. The minimal DFA accepting a regular language L does not contain C 1 -construction if and only if L can be accepted by an MM-1QFA with bounded error and with probability over 7 9. Proof. If the minimal DFA accepting a regular language L does not contain C 1 -construction, then, by Theorem 2 we obtain that L can be accepted by a GFA. Therefore, by Remar 2, L can be accepted by an MM-1QFA with bounded error and with probability over 7 9. On the other hand, if L is accepted by an MM-1QFA with bounded error and with probability over 7, then, with Remar 2 9 we now that L can be accepted by a GFA. By Theorem 2, the minimal DFA accepting L does not contain C 1 -construction. Next, we present a few examples to show that L(QFA ) is still a proper subset of all regular languages. Let us first show an example of regular language that can be accepted by an MM-1QFA but not by any multi-letter QFA. Example 1. The language a b can be accepted by an MM-1QFA [1] but it cannot be accepted by any -letter QFA. Indeed, we can describe the minimal DFA M for accepting a b by Fig. 5. In addition, from this figure we can find that there exists a C -construction for any 2, which is visualized by Fig. 6. Next we provide another example which demonstrates that there exist regular languages acceptable neither by MM- 1QFAs nor by multi-letter QFAs with bounded error. However, we need a result from [6].

3012 D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 Fig. 5. A state transition diagram of DFA M accepting a b. Fig. 6. A C -construction in Fig. 6. Fig. 7. Automaton G. Fig. 8. An F-construction in the minimal DFA G. Definition 5 ([6]). A DFA with state transition function δ is said to contain an F-construction if and only if there are non-empty words t, z Σ + and two distinct states q 1, q 2 Q such that δ (q 1, z) = δ (q 2, z) = q 2, δ (q 1, t) = q 1, δ (q 2, t) = q 2. Proposition 7 ([6]). A language L can be accepted by a multi-letter QFA with bounded error if and only if the minimal DFA of L does not contain any F-construction. Example 2. We use an example from [3]. Let L be the language consisting of all words that start with any number of letters a and after first letter b (if there is one) there is an odd number of letters a. The minimal DFA G accepting L is depicted by Fig. 7. As proved by Ambainis et al [3], L cannot be accepted by MM-1QFAs with bounded error. Indeed, L cannot be accepted by any multi-letter QFA, either. Because there exists an F-construction (Fig. 8) in the minimal DFA G (Fig. 7), we get the result. In conclusion, we can describe the relations between the families of languages accepted by MO-1QFAs, MM-1QFAs, and multi-letter QFAs, denoted by L(MO), L(MM), and L(QFA ), respectively. We recall that the language (a + b) b is accepted with no error by a 2-letter QFA but cannot be accepted by any MM-1QFA with bounded error, while a b is accepted by an MM-1QFA but cannot be accepted by any multi-letter QFA. Therefore, both L(MM)\L(QFA ) and L(QFA )\L(MM) hold. Furthermore, we have that L(MO) L(MM) L(QFA ), where may be proper. However, by Example 2, we have nown that L(MM) L(QFA ) still is a proper subset of all regular languages. 4. Determining the equivalence between multi-letter quantum finite automata Determining whether or not two one-way (probabilistic, quantum) finite automata and sequential machines are equivalent is of importance and has been well studied [27,34,26,24,25]. Concerning multi-letter QFAs, this issue is much more complicated and a new technique is needed. Here, we consider only the case of unary languages, i.e., the input alphabet having one element. Our goal is to deal with the decidability of equivalence of unary multi-letter QFAs. More specifically, for any given 1 - letter QFA A 1 and 2 -letter QFA A 2 over the same input alphabet Σ = {σ }, our purpose is to determine whether or not they are equivalent.

D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 3013 For a -letter QFA A = (Q, Q acc, ψ 0, Σ, µ), we recall the probability P A (x) for A accepting input string x = σ 1 σ 2 σ m and the definition of µ(x) as follows: { µ(λ 1 σ µ(x) = 1 )µ(λ 2 σ 1 σ 2 ) µ(λ m x), if m <, µ(λ 1 σ 1 )µ(λ 2 (4) σ 1 σ 2 ) µ(σ m +1 σ m +2 σ m ), if m, and then P A (x) = ψ 0 µ(x)p acc 2. We give the definition of equivalence between two multi-letter QFAs. Definition 6. A 1 -letter QFA A 1 and another 2 -letter QFA A 2 over the same input alphabet Σ are said to be equivalent (resp. t-equivalent) if P A1 (w) = P A2 (w) for any w Σ (resp. for any input string w with w t). Before we present a method for determining the equivalence between multi-letter QFAs over the same unary alphabet, we prove a useful lemma that is helpful to the main result. We recall the definition of tensor product of matrices [16]. For a 11 a 12 a 1n a 21 a 22 a 2n m n matrix A =.. and p q matrix B, their tensor product A B is an mp nq matrix defined as. a m1 a m2 a mn a 11 B a 12 B a 1n B a 21 B a 22 B a 2n B A B =... a m1 B a m2 B a mn B A basic property of tensor product is that, for any m n matrix A, p q matrix B, n o matrix C, and q r matrix D, (A B)(C D) = (AC) (BD). Now we present the crucial lemma. Lemma 8. Let {U 1, U 2,..., U } be a finite set of n n unitary matrices, and let M n 2 denote the linear space consisting of all n 2 n 2 complex square matrices. Denote H (i) = span{(u 1 U 2 U ) (U 1 U 2 U ),..., (U 1U 2 U i ) (U 1 U 2 (U i ) )} for i = 1, 2,..., where, for any subset A of M n 2, span A denotes the minimal subspace spanned by A, and denotes the conjugate operation. Then, there exists an i 0 n 4 such that H (i 0) for any t 0. = H (i 0+t) Proof. Let dim(s) denote the dimension of subspace S. Due to H (i) H (i+1) M n 2, for any i 1, we have 1 dim(h (1) ) dim(h (2) ) dim(h (n4 +1) ) n 4. Therefore, we obtain that there exists an i 0 n 4 such that H (i0) = H (i0+1). Next we prove by induction that Eq. (6) holds for t 0. First, we have nown that it holds for t = 0, 1. Suppose that it holds for t = i 1, i.e., H (i0) = H (i0+i). This implies that H (i0) = H (i0+1) = = H (i0+i). Our purpose is to show that it holds for t = i + 1, i.e., H (i0) = H (i0+i+1). Indeed, we have (U 1 U 2 U i 0+i+1 ) (U 1 U 2 (U i 0+i+1 ) ) = where (7) is due to the assumption H (i 0) = H (i 0+i). Therefore, (U 1 U 2 U i 0+i+1 ) (U 1 U 2 (U i 0+i+1 ) ) H (i0+1) = H (i0+i). = = [ (U 1 U 2 U i 0+i ) (U 1 U 2 (U i 0+i ) ) i 0 j=1 i 0 j=1 ] (U U ) c j [(U 1 U 2 U j ) (U 1 U 2 (U j ) )](U U ) (7) c j (U 1 U 2 U j+1 ) (U 1 U j+1 2 (U ) ) (8) Consequently, H (i0+i+1) = H (i0+i). Again, by the assumption of induction H (i0) = H (i0+i), we obtain that H (i0+i+1) = H (i0). Therefore, (6) holds for any t 0. (5) (6)

3014 D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 Now we are ready to present the main theorem regarding the equivalence of multi-letter QFAs. Theorem 9. For Σ = {σ }, a 1 -letter QFA A 1 = (Q 1, Q acc,1, ψ (1), Σ, µ 0 1) and another 2 -letter QFA A 2 = (Q 2, Q acc,2, ψ (2), Σ, µ 0 2) are equivalent if and only if they are (n 1 + n 2 ) 4 + 1-equivalent, where n i is the number of states of Q i, i = 1, 2, = max( 1, 2 ), with 1, 2 1. Proof. Let P acc,1 and P acc,2 denote the projections on the subspaces spanned by Q acc,1 and Q acc,2, respectively. For any string x Σ, we set µ(x) = µ 1 (x) µ 2 (x) and P acc = P acc,1 P acc,2, Q acc = Q acc,1 Q acc,2, where denotes the direct sum operation [ of ] any two matrices. More precisely, for any m 1 n 1 matrix A and m 2 n 2 matrix B, A B is defined as A 0 A B =, an (m 0 B 1 + m 2 ) (n 1 + n 2 ) matrix. In addition, we denote η 1 = ψ (1) 0 0 2 and η 2 = 0 2 ψ (2) 0, where 0 1 and 0 2 represent column zero vectors of n 1 and n 2 dimensions, respectively. Then, for any string x Σ, P η1 (x) = η 1 µ(x)p acc 2 (9) and P η2 (x) = η 2 µ(x)p acc 2. (10) Indeed, we further have that P η1 (x) = η 1 µ(x)p acc 2 = η 1 µ(x)p acc P Ď acc µ(x)ď η 1 = η 1 µ(x)p acc µ(x) Ď η 1 = ψ (1) 0 µ 1(x)P acc,1 µ 1 (x) Ď ψ (1) 0 = P A1 (x) (11) and P η2 (x) = η 2 µ(x)p acc 2 = η 2 µ(x)p acc P Ď acc µ(x)ď η 2 = η 2 µ(x)p acc µ(x) Ď η 2 = ψ (2) 0 µ 2(x)P acc,2 µ 2 (x) Ď ψ (2) 0 = P A2 (x). (12) Therefore, P A1 (x) = P A2 (x) holds if and only if P η1 (x) = P η2 (x) (13) for any string x Σ. On the other hand, we have that P η1 (x) = η 1 µ(x)p acc 2 = η 1 µ(x) p j 2 = η 1 µ(x) p j ( η 1 µ(x) p j ) = η 1 ( η 1 ) µ(x) (µ(x)) p j ( p j ) = η 1 ( η 1 ) µ(x) (µ(x)) p j ( p j ) (14)

D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 3015 and P η2 (x) = η 2 µ(x)p acc 2 = η 2 µ(x) p j 2 = η 2 µ(x) p j ( η 2 µ(x) p j ) = η 2 ( η 2 ) µ(x) (µ(x)) p j ( p j ) = η 2 ( η 2 ) µ(x) (µ(x)) Therefore, Eq. (13) holds if and only if η 1 ( η 1 ) µ(x) (µ(x)) for any string x Σ. Denote p j ( p j ). (15) p j ( p j ) = η 2 ( η 2 ) µ(x) (µ(x)) p j ( p j ) (16) D(x) = µ(x) µ(x) (17) where D(x) is an (n 1 + n 2 ) 2 (n 1 + n 2 ) 2 complex square matrix. Then the equivalence between A 1 and A 2 depends on whether or not the following equation holds for all string x Σ : η 1 ( η 1 ) D(x) p j ( p j ) = η 2 ( η 2 ) D(x) p j ( p j ) (18) Consider the linear space M n 2 consisting of all (n 1 +n 2 ) 2 (n 1 +n 2 ) 2 complex square matrices. It is clear that the dimension of M n 2 equals (n 1 + n 2 ) 4 = n 4. By D (i) we denote the subspace of M n 2 spanned by {D(x) : x Σ, x i}, where x denotes the length of x. Clearly, we have D (0) D (1) D (i) D (i+1). (19) Since the dimension of D (i) is not more than (n 1 + n 2 ) 4 for any i 1, there exists i 0 such that for any N i 0, D (i0) = D (N). In the rest of the proof, our purpose is to fix i 0. For the sae of convenience, we denote A i = µ 1 (Λ 1 i σ i ) for i = 1, 2,..., 1, and B j = µ 2 (Λ 2 j σ j ) for j = 1, 2,..., 2. Set = max( 1, 2 ). If 1 2, then we denote A i = A 1 for i = 1 + 1, 1 + 2,..., ; if 2 1, then we denote B j = B 2 for j = 2 + 1, 2 + 2,...,. In addition, we denote C i = A i B i for i = 1, 2,...,. Then C i is an n = n 1 + n 2 order unitary matrix for i = 1, 2,...,. According to the definition of µ(x) = µ 1 (x) µ 2 (x), we now that C 1 C 2 C i = µ(x) for x Σ and x = i. On the other hand, if i, then C 1 C 2 C i +1 = µ(x). Thus, D(x) = (C 1 C 2 C i ) (C 1 C 2 C ) i for x Σ and x = i ; and if i, then D(x) = (C 1 C 2 C i +1 ) (C 1 C i +1 2 (C ) ). We set E (i) = span{d(x) : x Σ, x + i}, i = 0, 1, 2,.... Then, by means of Lemma 8 it follows that, there exists i 0 n 4 1, such that E (i 0) = E (i 0+s) (20) for any s 0. Eq. (20) implies that, for any x Σ with x + i 0, D(x) can be linearly represented by some matrices in {D(y) : y + i 0 }. Therefore, if Eq. (18) holds for x n 4 + 1, then so does it for any x Σ. We have proved this theorem. Remar 3. We analyze the complexity of computation in Theorem 11. As in [34], we assume that all the inputs consist of complex numbers whose real and imaginary parts are rational numbers and that each arithmetic operation on rational numbers can be done in constant time. Still we denote n = n 1 +n 2. Note that in time O(in 4 ) we chec whether or not Eq. (18) holds for x Σ with x = i. Because the length of x to be checed in Eq. (18) is at most n 4 +, the time complexity for checing whether the two multi-letter QFAs are equivalent is O(n 3 (1+2+ +(n 4 +)), that is at most O(n 12 + 2 n 4 +n 8 ).

3016 D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 5. Concluding remars In this paper, we have considered several issues concerning multi-letter QFAs. Our technical contributions mainly contain the following two aspects: (1) we have shown that ( + 1)-letter QFAs are strictly more powerful than -letter QFAs, that is, ( + 1)-letter QFAs can accept some regular languages unacceptable by any -letter QFA, and some examples of regular languages unacceptable by multi-letter QFAs have been provided. We have nown that multi-letter QFAs are strictly more powerful than MO-1QFAs [26], but they are not comparable to MM-1QFAs [23,1] since the language a b can be accepted with bounded error by MM-1QFAs but cannot be accepted by multi-letter QFAs, and the language (a + b) b shows the opposite direction. Moreover, a b(a 2 ) a cannot be accepted by MM-1QFAs and by multi-letter QFAs with bounded error. (2) We have proved that a 1 -letter QFA A 1 and another 2 -letter QFA A 2 for accepting unary languages are equivalent if and only if they are (n 1 + n 2 ) 4 + 1-equivalent, and the time complexity of this computing method is O(n 12 + 2 n 4 + n 8 ), where n = n 1 + n 2, n 1 and n 2 are the numbers of states of A 1 and A 2, respectively, and = max( 1, 2 ). The method presented in the paper may be generalized to deal with more general cases. Another issue worthy of consideration is concerning the state complexity of multi-letter QFAs compared with the usual 1QFAs for accepting some languages (for example, unary regular languages [36,31]). Also, the power of measure-many multi-letter QFAs, as the relation between MM-1QFAs and MO-1QFAs, is worth being clarified. Whether or not measure-many multi-letter QFAs can recognize non-regular languages may also be considered in the future. Acnowledgements The authors are very grateful to the two referees and Professor Ohotin for their invaluable comments and suggestions that helped greatly to improve the quality of the original manuscript. Also, we than our group of quantum computing of SYSU for finding some mistaes in the original manuscript. References [1] A. Ambainis, R. Freivalds, One-way quantum finite automata: Strengths, weanesses and generalizations, in: Proceedings of the 39th Annual Symposium on Foundations of Computer Science, IEEE Computer Society Press, Palo Alfo, CA, USA, 1998, pp. 332 341. Also quant-ph/9802062, 1998. [2] M. Amano, K. Iwama, Undecidability on quantum finite automata, in: Proceedings of the 31st Annual ACM Symposium on Theory of Computing, Atlanta, GA, USA, 1999, pp. 368 375. [3] A. Ambainis, A. Kiusts, M. Valdats, On the class of languages recognizable by 1-way quantum finite automata, in: 8th International Symposium on Theoretical Aspects of Computer Science, STACS 2001, in: Lecture Notes in Computer Science, vol. 2010, Springer, Berlin, 2001, pp. 75 86. [4] A. Bertoni, C. Mereghetti, B. Palano, Quantum computing: 1-way quantum automata, in: Proceedings of the 9th International Conference on Developments in Language Theory, DLT 2003, in: Lecture Notes in Computer Science, vol. 2710, Springer, Berlin, 2003, pp. 1 20. [5] A. Brodsy, N. Pippenger, Characterizations of 1-way quantum finite automata, SIAM Journal on Computing 31 (2002) 1456 1478. Also quantph/9903014, 1999. [6] A. Belovs, A. Rosmanis, J. Smotrovs, Multi-letter reversible and quantum finite automata, in: Proceedings of the 13th International Conference on Developments in Language Theory, DLT 2007, Harrachov, Czech Republic, in: Lecture Notes in Computer Science, vol. 4588, Springer, Berlin, 2007, pp. 60 71. [7] E. Bernstein, U. Vazirani, Quantum complexity theory, SIAM Journal on Computing 26 (5) (1997) 1411 1473. [8] P. Benioff, The computer as a physical system: A microscopic quantum mechanical Hamiltonian model of computers as represented by Turing machines, Journal of Statistic Physics 22 (1980) 563 591. [9] K. Culi II, J. Karhumäi, On the equivalence problem for deterministic multitape automata and transducers, in: Proceedings of the 6th Annual Symposium on Theoretical Aspects of Computer Science, STACS 1989, Paderborn, Germany, February 16 18, 1989, pp. 468 479. [10] C.S. Calude, M.J. Dinneen, K. Svozil, Reflections on quantum computing, Complexity 6 (1) (2000) 35 37. [11] C.S. Calude, B. Pavlov, Coins, quantum measurements, and Turing s barrier, Quantum Information Processing 1 (1 2) (2002) 107 127. [12] D. Deutsch, Quantum theory, the Church Turing principle and the universal quantum computer, Proceedings of the Royal Society of London Series A 400 (1985) 97 117. [13] D. Deutsch, Quantum computational networs, Proceedings of the Royal Society of London Series A 400 (1985) 73 90. [14] R.P. Feynman, Simulating physics with computers, International Journal of Theoretical Physics 21 (1982) 467 488. [15] L.K. Grover, A fast quantum mechanical algorithm for database search, in: Proceedings of the 28th Annual ACM Symposium on Theory of Computing, Philadelphia, PA, USA, 1996, pp. 212 219. [16] J. Grusa, Quantum Computing, McGraw-Hill, London, 1999. [17] M. Hirvensalo, Quantum Computing, second edition, Springer, Berlin, 2004. [18] M. Hirvensalo, Improved undecidability results on the emptiness problem of probabilistic and quantum cut-point languages, in: Proceedings of the 33rd International Conference on Current Trends in Theory and Practice of Computer Science, SOFSEM 2007, Harrachov, Czech Republic, in: Lecture Notes in Computer Science, vol. 4362, Springer, Berlin, 2007, pp. 309 319. [19] J. Hromovič, One-way multihead deterministic finite automata, Acta Informatica 19 (1983) 377 384. [20] V. Halava, T. Harju, J. Karhumäi, Undecidability in ω-regular languages, Fundamenta Informaticae 73 (1 2) (2006) 119 125. [21] T. Harju, J. Karhumäi, Decidability of the multiplicity equivalence of multitape finite automata, in: Proceedings of the 22nd Annual ACM Symposium on Theory of Computing, Baltimore, MD, USA, 1990, pp. 477 481. [22] J.E. Hopcroft, J.D. Ullman, Introduction to Automata Theory, Language, and Computation, Addison-Wesley, Massachusetts, 1979. [23] A. Kondacs, J. Watrous, On the power of finite state automata, in: Proceedings of the 38th IEEE Annual Symposium on Foundations of Computer Science, Miami Beach, FL, USA, 1997, pp. 66 75. [24] L.Z. Li, D.W. Qiu, Determination of equivalence between quantum sequential machines, Theoretical Computer Science 358 (2006) 65 74. [25] L.Z. Li, D.W. Qiu, Determining the equivalence for one-way quantum finite automata, Theoretical Computer Science 403 (2008) 42 51. [26] C. Moore, J.P. Crutchfield, Quantum automata and quantum grammars, Theoretical Computer Science 237 (2000) 275 306. Also quant-ph/9707031, 1997. [27] A. Paz, Introduction to Probabilistic Automata, Academic Press, New Yor, 1971. [28] D.W. Qiu, Characterization of sequential quantum machines, International Journal of Theoretical Physics 41 (2002) 811 822. [29] D.W. Qiu, Automata theory based on quantum logic: Reversibilities and pushdown automata, Theoretical Computer Science 386 (2007) 38 56. [30] D.W. Qiu, L.Z. Li, An overview of quantum computation models: Quantum automata, Frontiers of Computer Science in China 2 (2) (2008) 193 207.

D. Qiu, S. Yu / Theoretical Computer Science 410 (2009) 3006 3017 3017 [31] in: G. Rozenberg, A. Salomaa (Eds.), Handboo of Formal Languages, vol. 1, Springer-Verlag, Berlin, 1997. [32] P.W. Shor, Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer, SIAM Journal on Computing 26 (5) (1997) 1484 1509. [33] A. Salomaa, Computation and Automata, Cambridge University Press, Cambridge, 1985. [34] W.G. Tzeng, A polynomial-time algorithm for the equivalence of probabilistic automata, SIAM Journal on Computing 21 (2) (1992) 216 227. [35] A.C. Yao, Quantum circuit complexity, in: Proceedings of the 34th IEEE Symposium on Foundations of Computer science, 1993, pp. 352 361. [36] S. Yu, Regular Languages, in: G. Rozenberg, A. Salomaa (Eds.), Handboo of Formal Languages, Springer-Verlag, Berlin, 1998, pp. 41 110.