Simplest Non-Regular Deterministic Context-Free Language
Abstract
We introduce a new notion of -simple problems for a class of decision problems (i.e. languages), w.r.t. a particular reduction. A problem is -simple if it can be reduced to each problem in . This can be viewed as a conceptual counterpart to -hard problems to which all problems in reduce. Our concrete example is the class of non-regular deterministic context-free languages (DCFL′), with a truth-table reduction by Mealy machines (which proves to be a preorder). The main technical result is a proof that the DCFL′ language is DCFL′-simple, which can thus be viewed as the simplest problem in the class DCFL′.
This result has already provided an application, to the computational model of neural networks 1ANN at the first level of analog neuron hierarchy. This model was proven not to recognize , by using a specialized technical argument that can hardly be generalized to other languages in DCFL′. By the result that is DCFL′-simple, w.r.t. the reduction that can be implemented by 1ANN, we immediately obtain that 1ANN cannot accept any language in DCFL′.
It thus seems worthwhile to explore if looking for -simple problems in other classes under suitable reductions could provide effective tools for expanding the lower-bound results known for single problems to the whole classes of problems.
Keywords: deterministic context-free language, truth-table reduction, Mealy automaton, pushdown automaton
1 Introduction
We introduce a new notion of -simple problems for a class of decision problems (i.e. languages). A problem is -simple if it can be reduced to each problem in ; if this problem is, moreover, in , it can be viewed as a simplest problem in . The -simple problems are thus a conceptual counterpart to the common -hard problems (like, e.g., NP-hard problems) to which conversely any problem in reduces. These definitions (of -simple and -hard problems) are parametrized by a chosen reduction that does not have a higher computational complexity than the class itself. Therefore, it may be said that if a -hard problem has a (computationally) “easy” solution, then each problem in has an “easy” solution. On the other hand, if we prove that a -simple problem is not “easy”, in particular that it cannot be solved by machines of a type that can implement the respective reduction, then all problems in are not “easy”, that is, are not solvable by ; this extends a lower-bound result for one problem to the whole class of problems.
In this paper, we consider to be the class of non-regular deterministic context-free languages, which we denote by DCFL′; we thus have DCFL′ = DCFL REG (where REG denotes the class of regular languages). We use a truth-table reduction by Mealy machines (which is motivated below). Hence a DCFL′-simple problem is a language (over an alphabet ) that can be reduced to each DCFL′ language by a Mealy machine with an oracle , denoted . More precisely, the finite-state transducer transforms a given input word to a prefix of queries for the oracle . In addition, each state of is associated with a finite tuple of query suffixes from , and with a truth table . After reads an input word (translating it to ), by which it enters a state , for each it queries whether or not the string is in (or, equivalently, whether or not belongs to the quotient ), and aggregates the answers by the truth table for deciding if is accepted.
This truth-table reduction by Mealy machines proves to be a preorder, denoted as . The main technical result of this paper is that the DCFL′ language (over the binary alphabet ) is DCFL′-simple, since for each language in DCFL′. The class DCFLS of DCFL′-simple languages comprises REG and is a strict subclass of DCFL; e.g., the DCFL′ language over the alphabet proves to be not DCFL′-simple. The closure properties of DCFLS are similar to that of DCFL as the class DCFLS is closed under complement and intersection with regular languages, while being not closed under concatenation, intersection, and union.
The above definition of DCFL′-simple problems has originally been motivated by the analysis of the computational power of neural network (NN) models which is known to depend on the (descriptive) complexity of their weight parameters [8, 11]. The so-called analog neuron hierarchy [9] of binary-state NNs with increasing number of extra analog-state neurons, denoted as ANN for , has been introduced for studying NNs with realistic weights between integers (finite automata) and rational numbers (Turing machines). We use the notation ANN also for the class of languages accepted by ANNs, which can clearly be distinguished by the context. The separation 1ANN 2ANN has been witnessed by the DCFL′ language 2ANN 1ANN. The proof of 1ANN is rather technical (based on the Bolzano-Weierstrass theorem) which could hardly be generalized to other DCFL′ languages, while it was conjectured that 1ANN for all DCFL′ languages , that is, (implying ). An idea how to prove this conjecture is to show that 1ANN is in some sense the simplest problem in the class DCFL′, namely, to reduce to any DCFL′ language by using a reduction that can be carried out by 1ANNs, which are at least as powerful as finite automata. This would imply that cannot be accepted by any 1ANN since it is at least as hard as that has been proven not to be recognized by 1ANNs.
The idea why should serve as the simplest language in the class DCFL′ comes from the fact that any reduced context-free grammar generating a non-regular language is self-embedding [3, Theorem 4.10]. This means that there is a so-called self-embedding nonterminal admitting the derivation for some non-empty strings . Since is reduced, there are strings such that and where is the start nonterminal in , which implies for every . It is thus straightforward to suggest to reduce an input word where , to the string (while the inputs outside are mapped onto some fixed string outside ) since entails .
However, the suggested (one-one) reduction from to is not consistent because does not necessarily imply . For example, consider the DCFL′ language over the binary alphabet for which there are no words such that would ensure . Nevertheless, we can pick two inputs and instead of one, that is, , , and ( denoting the empty string), which satisfy iff iff and . It turns out that this can be generalized to any DCFL′ language. Namely, we prove in this paper that for DCFL′ language over any alphabet , there are non-empty words and a language , where is the complement of , such that iff and .
Therefore, the simple many-one (in fact, one-one) reduction from with one query to the oracle is replaced by a truth-table reduction, that is, by a special Turing reduction in which all its finitely many (in our case two) oracle queries are presented at the same time and there is a Boolean function (a truth table) which, when given the answers to the queries, produces the final answer of the reduction. This truth-table reduction from to can be implemented by a deterministic finite-state transducer (a Mealy machine) with the oracle : It transforms the input where (the inputs outside are rejected), to the output and carries out two queries to that arise by concatenation of this output with two fixed suffixes and ; hence the queries are and . The truth table is defined so that the input is accepted by iff the two answers to these queries are distinct and at same time, the first answer is negative in the case , and positive in the case , which is equivalent to .
It follows that the DCFL′ language is DCFL′-simple under the truth-table reduction by Mealy machines. Since this reduction can be implemented by 1ANNs, we achieve the desired stronger separation in the analog neuron hierarchy [10]. This result constitutes a non-trivial application of the proposed concept of DCFL′-simple problem. Moreover, if we could generalize the result to (nondeterministic) context-free languages (CFL), e.g. by proving that some DCFL′ language is CFL′-simple (where CFL′ CFL REG), which would imply that is CFL′-simple by the transitivity of reduction, then we would achieve even stronger separation . We note the interesting fact that cannot be CSL′-simple (under our reduction), since 1ANN accepts some context-sensitive languages outside CFL [9].
In general, if we show that some -simple problem under a given reduction cannot be computed by a computational model that implements this reduction, then all problems in the class are not solvable by either. The notion of -simple problems can thus be useful for expanding known (e.g. technical) lower-bound results for individual problems to the whole classes of problems at once, as it was the case of the DCFL′-simple problem , expanding to . It seems worthwhile to explore if looking for -simple problems in other complexity classes could provide effective tools for strengthening known lower bounds.
We remark that the hardest context-free language by Greibach [2] can be viewed as CFL-hard under a special type of our reduction . Related line of study concerns the types of reductions used in finite or pushdown automata with oracle. For example, nondeterministic finite automata with oracle complying with many-one restriction have been applied to establishing oracle hierarchies over the context-free languages [7]. For the same purpose, oracle pushdown automata have been used for many-one, truth-table, and Turing reducibilities, respectively, inducing the underlying definitions also to oracle nondeterministic finite automata [13]. In addition, nondeterministic finite automata whose oracle queries are completed by the prefix of an input word that has been read so far and the remaining suffix, have been employed in defining a polynomial-size oracle hierarchy [1].
In the preliminary study [12], some considerations about the simplest DCFL′ language have appeared, yet without formal definitions of DCFL′-simple problems, that included only sketches of incomplete proofs of weaker results based on the representation of DCFL by so-called deterministic monotonic restarting automata [5], which have initiated investigations of non-regularity degrees in DCFL [6].
In this paper we achieve a complete argument for to be a DCFL′-simple problem, within the framework of deterministic pushdown automata (DPDA) by using some ideas on regularity of pushdown processes from [4]. We now give an informal overview of the proof. Given a DPDA recognizing a non-regular language , it is easy to realize that some computations of (from the initial configuration) must be reaching configurations where the stack is arbitrarily large while it can be (almost) erased afterwards. Hence the existence of words such that for all is obvious. However, we aim to guarantee that for all the equality holds if, and only if, and , where is either the language or its complement. This is not so straightforward but it is confirmed by our detailed analysis (in section 3). We study the computation of on an infinite word that visits infinitely many pairwise non-equivalent configurations. We use a natural congruence property of language equivalence on the set of configurations, and avoid some tedious technical details by a particular use of Ramsey’s theorem. This allows us to extract the required tuple from the mentioned infinite computation. We note that determinism of is essential in the presented proof; we leave open if it can be relaxed to show that is even CFL′-simple.
The rest of the paper is organized as follows. In section 2 we recall basic definitions and notation regarding DPDA and Mealy machines, introduce the novel concept of DCFL′-simple problems under truth-table reduction by Mealy machines and show some simple properties of the class DCFLS of DCFL′-simple problems. In section 3 we present the proof of the main technical result which shows that is DCFL′-simple. Finally, we summarize the results and list some open problems in section 4.
2 DCFL′-Simple Problem Under Truth-Table Mealy Reduction
In this section we define the truth-table reduction by Mealy machines, introduce the notion of DCFL′-simple problems, show their basic properties, and formulate the main technical result (theorem 1). But first we recall standard definitions of pushdown automata.
A pushdown automaton (PDA) is a tuple where is a finite set of states including the start state and the set of accepting states, while the finite sets and represent the input and stack alphabets, respectively, with the initial stack symbol . In addition, the set contains finitely many transition rules with the meaning that in state , on the input (recall denotes the empty string), and with as the topmost stack symbol may read , change the state to , and pop , replacing it by pushing .
By a configuration of we mean , and we define relations for on : each rule in induces for all ; these relations are naturally extended to for . For a configuration we define , and is the language accepted by . A PDA is deterministic (a DPDA) if there is at most one rule for each tuple , , ; moreover, if there is a rule , then there is no rule for . We also use the standard assumption that all -steps are popping, that is, in each rule in we have .
The languages accepted by (deterministic) pushdown automata constitute the class of (deterministic) context-free languages; the classes are denoted by DCFL and CFL, respectively, whereas DCFL′ DCFL REG.
In the following theorem we formulate the main technical result: any language in DCFL′ includes a certain “projection” of the language , which means that is in some sense the simplest language in the class DCFL′. The theorem, whose proof will be presented in section 3, thus provides an interesting property of DCFL′.
Theorem 1.
Let be a non-regular deterministic context-free language over an alphabet . There exist non-empty words and a language (where is the complement of ) such that for all and we have
(1) |
In order to formalize the DCFL′-simple problems, we now define a Mealy machine with an oracle: it is a tuple where is a finite set of states including the start state , and the finite sets and represent the input and output (oracle) alphabets, respectively. Moreover, is a (partial) state-transition function which extends to input strings as where for every , while for all , , . Similarly, is an output function which extends to input strings as where for all , and for all , , . In addition, for each , the tuple of strings in contains query suffixes, while is a truth table that aggregates the answers to the oracle queries.
The above Mealy machine starts in the start state and operates as a deterministic finite-state transducer that transforms an input word to the output string written to a so-called oracle tape. The oracle tape is a semi-infinite, write-only tape which is empty at the beginning and its contents are only extended in the course of computation by appending the strings to the right. Namely, given a current state and an input symbol , the machine moves to the next state and writes the string to the oracle tape, if is defined; otherwise rejects the input. After reading the whole input word , the machine is in the state , while the oracle tape contains the output .
Finally, the Mealy machine , equipped with an oracle , in this case denoted , queries the oracle whether belongs to the (right) quotient , for each suffix in , and the answers are aggregated by the truth table . Thus, the oracle Mealy machine accepts the input word iff
where and is the characteristic function of , that is, if , and if . The language accepted by the machine is defined as is accepted by .111Note that the described protocol works also for non-prefix-free languages since for any input prefix that has been read so far, the output value from the truth table determines whether the oracle Mealy machine is in an “accepting” state, deciding about this prefix analogously as a deterministic finite automaton. The truth-table reduction only requires that the given oracle answers do not influence further computation when subsequent input symbols are read.
We say that is truth-table reducible to by a Mealy machine, which is denoted as , if for some Mealy machine running with the oracle . The following lemma shows that we can chain these reductions together since the relation is a preorder.
Lemma 2.
The relation is reflexive and transitive.
Proof: The relation is reflexive since for the oracle Mealy machine where and for every , , and is the identity.
Now we show that the relation is transitive. Let and which means and for some oracle Mealy machines and , respectively. We will construct the oracle Mealy machine such that which implies the transitivity . We define with , and for every and , which ensures for every . For each state in , we define the tuple of query suffixes from ,
where and are the query suffixes associated with and for , respectively, and the truth table aggregates the answers to the corresponding oracle queries, which ensures .
∎
We say that a (decision) problem is DCFL′-simple if for every non-regular deterministic context-free language . It follows from theorem 1 that the DCFL′ language is an example of a DCFL′-simple problem. In addition, we denote by DCFLS the class of DCFL′-simple problems and formulate its basic properties.
Corollary 3 (of theorem 1).
The non-regular deterministic context-free language is DCFL′-simple.
Proof:
Let be any DCFL′ language. According
to theorem 1, there are and
such that condition (1)
holds for .
We define the Mealy machine with the oracle , as , , , , , ,
,
, ,
and where iff .
It is easy to verify that , which
implies . Hence, is DCFL′-simple.
∎
Proposition 4.
-
1.
.
-
2.
, and .
-
3.
The class DCFLS is closed under complement and intersection with regular languages.
-
4.
The class DCFLS is not closed under concatenation, intersection and union.
Proof: [Sketch.]
1.
For any regular language , consider a Mealy machine with the DCFL′-simple oracle , that simulates a deterministic finite automaton recognizing , while its constant truth tables produce 1 iff associated with the accept states. Hence, which means is DCFL′-simple according to lemma 2 and corollary 3 which also implies .
2. We first observe that . Let DCFLS be any DCFL′-simple language which ensures by an oracle Mealy machine . The machine can be simulated by a DPDA which extends a suitable DPDA (e.g. with no -transitions) accepting , so that the finite control of implements the finite-state transducer whose output is presented online as an input to . Moreover, for each state of , the finite control of evaluates the truth table which aggregates the answers to the queries with suffixes associated with , by inspecting at most constant number of topmost stack symbols. Hence DCFL.
In order to show that , we prove that the DCFL over the alphabet is not DCFL′-simple. For the sake of contradiction, suppose that by a Mealy machine with the oracle , which means . Consider all the possible prefixes of inputs presented to that have the length . These strings can bring into a finite number of distinct states while the length of outputs written to the oracle tape is bounded by . For outside , the acceptance of words where , depends only on the truth values associated with the states from the finite set , due to for any . On the other hand, the number of distinct outputs in is bounded by . This means that for a sufficiently large , there must be two distinct prefixes such that and in , which results in the contradiction .
3. The class DCFLS is closed under complement since the truth tables can be negated. Furthermore, any oracle Mealy machine be can modified so that it simulates another given finite automaton in parallel and is forced to reject if this automaton rejects, which shows DCFLS to be closed under intersection with regular languages.
4.
Observe that is not DCFL′-simple under truth-table reduction. In addition, and are DCFL′-simple while is not context-free. The proof for union follows from 3 and De Morgan’s law.
∎
3 Proof of the Main Result (Theorem 1)
Theorem 1 follows from the (more specific) next lemma that we prove in this section.
By we denote the set , and by the set (for ).
Lemma 5.
Let be a DPDA where is non-regular (hence belongs to DCFL′). There are , , , , , such that the following four conditions hold:
-
1.
and ,
which entails the infinite (stack increasing) computation(2) -
2.
;
-
3.
,
hence for all and ;
-
4.
one of the following cases is valid (depending on whether or ):
-
(a)
iff (for all ), or iff (for all );
-
(b)
iff (for all ), or iff (for all ).
-
(a)
We note that (for each ); hence iff (since is nonempty). Theorem 1 indeed follows from the lemma: there is such that either iff (for all ), or iff (for all ). (In theorem 1 we also stated that is nonempty. If here, then we simply take and as the new , respectively.)
Proof of Lemma 5
In the rest of this section we provide a proof of lemma 5, assuming a fixed DPDA where is non-regular. The proof structure is visible from the auxiliary claims that we state and prove on the way.
Convention. W.l.o.g. we assume that always reads the whole input from . This can be accomplished in the standard way, by adding a special bottom-of-stack symbol and a (non-accepting) fail-state. (Each empty-stack configuration becomes , and each originally stuck computation enters the fail-state where it loops. We also recall that all -steps are popping, and thus infinite -sequences are impossible.) Hence for any infinite word in there is the unique infinite computation of starting in ; it stepwise reads the whole infinite word .
The left quotient of by is the set ; concatenation has priority over , hence . (The next claim is valid for any non-regular .)
Claim 6.
We can fix an infinite word in () such that for all .
Proof:
Let us consider the labelled transition system
where
and .
(We recall that entails .)
Since is non-regular, the set of states
reachable from in is
infinite. The out-degree of states in is finite
(in fact, bounded by ), hence an application of
König’s lemma yields
an infinite acyclic path
.
∎
We call a configuration of unstable if and contains a rule (we recall that -steps are only popping); otherwise is stable. Since is a deterministic PDA, for each unstable we can soundly define the stable successor of as the unique stable configuration where ( being a suffix of ). The path might (not) go via an accepting state (in ), hence or . (We note that the configurations in the computation (2) that start with are necessarily stable.)
Claim 7.
Each configuration is visited at most twice by
the computation of from on that is fixed by 6. | (3) |
Proof: The computation (3) is infinite, stepwise reading the whole word , and it can be presented as
(for )
where each is stable; each segment starts with a (visible) -step that is followed by a (maybe empty) sequence of (popping) -steps via unstable configurations. Since such an -sequence might go through an accepting state, we can have for though ; in this case contains precisely one of the words and , and the languages and differ just on . Nevertheless, this reasoning entails that we cannot have for pairwise different .
Since each segment
visits any unstable configuration at most once and
is the stable successor for all unstable
configurations in the segment, we deduce that also each unstable
configuration can be visited at most twice
in the computation (3).
∎
Claim 8.
The computation (3) on can be “stair-factorized”, that is, written
(4) |
so that for each we have and where is a nonempty suffix of the right-hand side of a rule in (i.e., a nonempty suffix of in a rule ).
Proof: We consider the computation (3), and call a stable configuration a level, with position , if and all configurations visited by the computation after have the stack longer than ; we note that each level has a unique position . Since each configuration is visited at most twice in (3), the set of levels is infinite, with elements , , , where . The computation (3) can thus be presented as
where , and for , putting .
Each segment
between two neighbouring levels can
be obviously written as
where is a rule in , both
and are nonempty, , and
.
Hence the validity of the claim is clear. ∎
We define the natural equivalence relation on the set of configurations of : we put if .
We fix the presentation (4), calling the level-configurations (for all ). Since we have , there cannot be three level-configurations in the same -class (i.e., in the same equivalence class w.r.t. ). Hence any infinite set of level-configurations represents infinitely many -classes. Now we show a congruence-property that might enable to shorten a level-configuration while keeping its -class. We use the notation (the “down-states” of ), putting
for some .
Claim 9.
If for each , then .
Proof:
Let us consider . If , then
for all .
If and
there is no prefix of such
that ,
then for all .
If and where
(necessarily for some ),
then iff . Hence the claim is clear.
∎
The next claim is an immediate corollary.
Claim 10.
Any computation where (), , and for each can be shortened to where .
The -th level-configuration in (4) is reached by the computation . It can happen that there are , such that and for all . In this case we can shorten the computation as in 10, where corresponds to the omitted . The resulting shorter computation might be possible to be repeatedly shortened further (if it can be presented so that the conditions of 10 are satisfied). Now for each we fix a (stair-factorized) computation
(5) |
that has arisen by a maximal sequence of the above shortenings of the prefix
of (4).
Hence
,
,
is a
subsequence of
,
and
.
Claim 11.
For each there is such that (where is from (5)).
Proof:
As already discussed,
the set of level-configurations represents infinitely many
-classes. The last configurations of
computations (5) represent the same
infinite set of -classes, and their lengths
thus cannot
be bounded;
since the lengths of all are bounded (they are
shorter than the longest right-hand sides of the rules in
), the claim is clear.
∎
Now we come to a crucial claim in our proof of lemma 5. Besides the notation we also introduce (the by--reached down-states of ), by putting
.
Hence , and (due to the determinism of the DPDA ).
We recall that means . To handle the special case of the empty word , we also define a (much) coarser equivalence : we put if either belongs to both and , or belongs to none of them.
Claim 12.
There is a constant determined by the DPDA such that for all where the final configuration in (5) can be written as
where the following conditions hold:
-
1.
where and
(and , ); -
2.
the sets and are equal, further being denoted by ;
-
3.
for each , if , then (and );
-
4.
each belongs to for some self-containing , where is self-containing if ;
-
5.
there is a state for which and .
Proof: We fix some with larger than a constant determined by as described below (there are such by 11). For convenience we put , , and , hence the final configuration in (5) is . We view the prefixes
as the vertices of a complete graph with coloured edges.
For , where , , and , , the edge between the vertices and has the following tuple as its colour:
where and (and are taken from (5)).
Since the set of colours is bounded (by a constant determined by ), Ramsey’s theorem yields a bound guaranteeing that there is a monochromatic clique of size among the vertices , , , , . (We have soundly chosen so that is bigger than .) We fix such a monochromatic clique MC, denoting its vertices as
, , , and its colour as .
This is sound, since the fact that both edges and have the same colour entails that the first component in this colour is the same as the second component, and the third component is the same as the fourth component.
We now show that the conditions – are satisfied for the presentation of as , where for the respective .
Conditions and are trivial (due to the colour C).
Condition 3: Let and (hence also ). Then (since MC is monochromatic). This entails , hence , which in turn entails .
Condition : We first note a general fact: . Since , for each there is thus such that . We also have the following “transitivity”: if , , and , then (since MC is monochromatic). For any there is clearly a “chain” where , for all , and for some . By the above transitivity, is self-containing ( and thus ) and (hence ).
Condition :
For any three configurations at least two belong to the same
-class.
Since the edges among the vertices
, ,
have the same in their colour C, we get
that , and thus also
for all such that
.
Now if for all we had
(which includes the case ),
then we would get a contradiction with our choice
of (5) since it could have been shortened as in
10.
∎
Now we are already close to lemma 5:
Claim 13.
There are , , , , , such that , , , , and
-
•
either and for all ,
-
•
or and for all .
Proof: We fix one guaranteed by 12 (satisfying the respective conditions –). There are , , , , , , such that
, , , , and and differ on
(i.e., .
(Indeed: The respective computation (5) can be written
where and are
nonempty.
The claimed and [nonempty] are guaranteed by
in
12, and is a respective self-containing
state from .
Since and ,
we get
, where
.
We also have , since
otherwise , , and we
could not have
and .)
Since , we can fix such that . Hence the languages and differ on ; more generally, and differ on for all . Now we aim to find out for which we have .
We recall that ; hence . Since , we get that for all (by induction). We now distinguish two cases:
-
1.
For each prefix of and each we have: if , then .
-
2.
There are a prefix of , , and such that and .
In the case we clearly have either or (here plays no role). In the case we recall that entails that for all (since by in 12). Hence we have either or .
Since and
differ on , we deduce that there is
such that
either
and
for all , or
and
for all .
Hence for we have
either
and
for all
,
or
and
for all .
Since for we have
, the claim is proven.
∎
Claim 13 is a weaker version of lemma 5; it shows that there is such that and for . To handle the case , we have to find out for which we have . We thus look at the computation from on the infinite word (recalling our convention that this computation is infinite, stepwise reading the word ), and use the obvious fact that after a prefix this computation becomes “periodic” (either cycling among finitely many configurations, or increasing the stack forever).
Claim 14.
For any configuration and words there are numbers and (“period”) such that for all the remainder determines whether or not .
Proof: We assume (otherwise the claim is trivial). For the infinite computation from on there are obviously , , , and such that the computation can be written where . (We have if the computation visits only finitely many configurations, and otherwise we consider the stair-factorization of the computation.)
For each we put , and we have two possible cases:
-
1.
There is such that for all performing from does not reach at the bottom.
-
2.
There are , a prefix of , , and such that and .
In the case either for all
, or for all
.
In the case , for each we have
where if
.
Hence
for each , the (non)membership of in is determined by .
The claim is thus clear.
∎
4 Conclusion and Open Problems
In this paper, we have introduced a new notion of the -simple problem that reduces to each problem in , being thus a conceptual counterpart to the -hard problem to which each problem in reduces. We have illustrated this concept on the definition of the DCFL′-simple problem that reduces to each DCFL′ language under the truth-table reduction by Mealy machines. We have proven that the DCFL′ language is DCFL′-simple, and thus represents the simplest languages in the class DCFL′. This result finds its application in expanding the known lower bound for , namely that cannot be recognized by the neural network model 1ANN, to all DCFL′ languages. Moreover, the class DCFLS of DCFL′-simple problems containing the regular languages is a strict subclass of DCFL and has similar closure properties as DCFL.
We note that the hardest context-free language by Greibach [2], where each in CFL is an inverse homomorphic image of or , can be viewed as CFL-hard w.r.t. a many-one reduction based on Mealy machines realizing the respective homomorphisms. Our aims in the definition of DCFL′-simple problems cannot be achieved by such a many-one reduction, hence we have generalized it to a truth-table reduction. We can alternatively consider a general Turing reduction that is implemented by a Mealy machine which queries the oracle at special query states, each associated with a corresponding query suffix, while its next transition from the query state depends on the given oracle answer. The oracle Mealy machine then accepts an input word if it reaches an accept state after reading the input. The language proves to be DCFL′-simple under this Turing reduction allowing for an unbounded number of online oracle queries; this can be shown by 13 (a weaker version of lemma 5).
It is natural to try extending our result to non-regular nondeterministic (or at least unambiguous) context-free languages, by possibly showing that is CFL′-simple. Another important challenge for further research is looking for -simple problems for other complexity classes and suitable reductions. This could provide an effective tool for strengthening lower-bounds results known for single problems to the whole classes of problems, which deserves a deeper study.
Acknowledgements
Presented research has been partially supported by the Czech Science Foundation, grant GA19-05704S, and by the institutional support RVO: 67985807 (J. Šíma). J. Šíma also thanks Martin Plátek for his intensive collaboration at the first stages of this research.
References
- [1] Anabtawi, M., Hassan, S., Kapoutsis, C.A., Zakzok, M.: An oracle hierarchy for small one-way finite automata. In: Proceedings of LATA 2019. pp. 57–69. LNCS 11417, Springer (2019). https://doi.org/10.1007/978-3-030-13435-8_4
- [2] Greibach, S.A.: The hardest context-free language. SIAM J. Comput. 2(4), 304–310 (1973). https://doi.org/10.1137/0202025
- [3] Hopcroft, J.E., Ullman, J.D.: Formal languages and their relation to automata. Addison-Wesley (1969), https://www.worldcat.org/oclc/00005012
- [4] Jančar, P.: Deciding semantic finiteness of pushdown processes and first-order grammars w.r.t. bisimulation equivalence. J. Comput. Syst. Sci. 109, 22–44 (2020). https://doi.org/10.1016/j.jcss.2019.10.002
- [5] Jančar, P., Mráz, F., Plátek, M., Vogel, J.: Restarting automata. In: Proceedings of FCT 1995. pp. 283–292. LNCS 965, Springer (1995). https://doi.org/10.1007/3-540-60249-6_60
- [6] Mráz, F., Pardubská, D., Plátek, M., Šíma, J.: Pumping deterministic monotone restarting automata and DCFL. In: Proceedings of ITAT 2020. pp. 51–58. CEUR Workshop Proceedings 2718 (2020), http://ceur-ws.org/Vol-2718/paper13.pdf
- [7] Reinhardt, K.: Hierarchies over the context-free languages. In: Proceedings of IMYCS 1990. pp. 214–224. LNCS 464, Springer (1990). https://doi.org/10.1007/3-540-53414-8_44
- [8] Siegelmann, H.T.: Neural networks and analog computation – Beyond the Turing limit. Birkhäuser (1999)
- [9] Šíma, J.: Analog neuron hierarchy. Neural Netw. 128, 199–215 (2020). https://doi.org/10.1016/j.neunet.2020.05.006
- [10] Šíma, J.: Stronger separation of analog neuron hierarchy by deterministic context-free languages (2021), arXiv:2102.01633 (submitted to a journal)
- [11] Šíma, J., Orponen, P.: General-purpose computation with neural networks: A survey of complexity theoretic results. Neural Comput. 15(12), 2727–2778 (2003). https://doi.org/10.1162/089976603322518731
- [12] Šíma, J., Plátek, M.: One analog neuron cannot recognize deterministic context-free languages. In: Proceedings of ICONIP 2019, Part III. pp. 77–89. LNCS 11955, Springer (2019). https://doi.org/10.1007/978-3-030-36718-3_7
- [13] Yamakami, T.: Oracle pushdown automata, nondeterministic reducibilities, and the hierarchy over the family of context-free languages. In: Proceedings of SOFSEM 2014. pp. 514–525. LNCS 8327, Springer (2014). https://doi.org/10.1007/978-3-319-04298-5_45, (full version arXiv:1303.1717)