Regular expressions Lexical Analysis - regex

Why repeated strings such as
[wcw|w is a string of a's and b's]
cannot be denoted by regular expressions?
pls. give me detailed answer as i m new to lexical analysis.
Thanks ...

Regular expressions in their original form describe regular languages/grammars. Those cannot contain nested structures as those languages can be described by a simple finite state machine. Simplified you can picture that as if each word of the language grows strictly from left to right (or right to left), where repeating structures have to be explicitly defined and are static.
What this means is, that no information whatsoever from previous states can be carried over to later states (a few characters further in the input). So if you have your symbol w you can't specify that the input must have exactly the same string w later in the sequence. Similarly you can't ensure that each opening paranthesis needs a closin paren as well (so regular expressions themselves are not even a regular language and thus cannot be described by regular expressions :-)).
In theoretical computer science we worked with a very restricted set of regex operators, basically only consisting of sequence, alternative (|) and repetition (*), everything else can be described with those operations.
However, usually regex engines allow grouping of certain sub-patterns into matches which can then be referenced or extracted later. Some engines even allow to use such a backreference in the search expression string itself, thereby allowing the expression to describe more than just a regular language. If I remember correctly such use of backreferences can even yield languages that are not context-free.
Additional pointers:
This StackOverflow question
Wikipedia

It can be, you just can't assure that it's the same string of "a"s and "b"s because there's no way to retain the information acquired in traversing the first half for use in traversing the second.

Related

Counterpart of regular expressions for parsing nested strucures

Regular expressions are a standard tool used for parsing strings across many languages. However their scope of applicability is limited. Regular expressions can only match a list. There is no way to describe arbitrary deep nested structures using regular expressions. Question: what is a technology/framework as widely used/spread and as standatd as regular expessions are that can match tree structures (produce AST).
Regular expressions describe a finite-state automaton.
Since the late 1960's, the "bread and butter" of parsing (though not necessarily the "state of the art") has been push-down automata generated by parser generators according to "LR" algorithms like LALR(1).
The connection to regular expressions is this: the parsing machine does in fact use rules very similar to regular expressions in order to recognize viable prefixes. The "shift" state transitions among the "core LR(0) items" constitute a finite automaton, and can be described by a regular expression. The recursion is is handled thanks to the semantic action of pushing symbols onto a stack when doing the "shifts", and removing them ("reducing"). Reductions rewrite a portion of the stack, and perform a "goto" to another state. This type of goto, together with the stack, is absent in the regular expression automaton.
Parse Expression Grammars are also related to regular expressions. Regular expressions themselves can be endowed with recursion. Firstly, we can take pieces of regular expressions and give them names, and then construct bigger regular expressions by writing expressions which invoke these names. (Such as feature is found in the lex tool where you can define a named expressions like letters [A-Za-z]+ and refer to it later as {letters}. Now suppose you allow circular references, like letters [A-Za-z]{letters}?. You now have recursion; the only problem is to adjust the model in order to implement it.
Implementations of so-called "regular expressions" in various modern languages and environments in fact support recursion. Perl-compatible regular expressions (PCRE) support it, for instance.
Expressions that feature recursion or backreferencing are not handled by the classic NFA compilation route (possibly converted to a DFA); they cannot be.
How the above letters recursion can be handled is with actual recursion. The ? operator can be implemented as a function which tries to match its respective argument object. If it succeeds, then it consumes whatever it has matched, otherwise it consumes nothing. That is to say, the regular expression can be converted to a syntax tree, and interpreted "as is" rather than compiled to a state machine (or trivially compiled to functions corresponding to the nodes of the tree), and such interpretation can naturally handle recursion. The interpretation then constitutes, effectively, a syntax-directed recursive-descent parser. (Note how I avoided left recursion in defining letters to make that example compatible with this approach).
Example: parenthesis-matching regex:
par-match := ({par-match})|
This gets compiled to a tree:
branch-op <-- "par-match" name points at this node
/ \
catenate-op <empty>
/ \
"(" catenate-op
/ \
{par-match} ")"
This can then converted to a recursive descent parser, or interpreted directly.
Pattern matching starts by invoking the top-level "branch-op". This operator simply tries all of the alternatives. Suppose the input is empty. Then the left alternative will fail: it demands an open parenthesis. So then the right alternative will succeed: empty matches empty. (The operators either "fail" or indicate "success" and consume input.)
But suppose your input is (()). The left catenate-op will in turn invoke its left subtree, which matches and consumes the left parenthesis, leaving ()). It will then invoke its right subtree, another catenate-op. This catenate-op matches its left subtree, which triggers recursion into the top level via the named par-match references. That recursion will match and consume (), leaving ). The catenate-op then invokes its right subtree which matches ). Control returns up to branch-op. (Though the left side of branch-op matched something, branch-op must still try the other alternative; more than one branch can match, and some can match longer than others.)
This is closely related to Parsing Expression Grammars work.
Practically speaking, the recursive definition could be encoded into the regex syntax somehow. Say we invent some new operator like (?name:definition) which means "match definition which is allowed to contain invocations of itself via name. The invocation syntax could be (*name), so that we can write the par-match example as (?par-match:\((*par-match)\)|). The combinations (? and (* are invalid under "classic" regex syntax and so we can use them for extension.
As a final note, regexes correspond to grammars. That is the fundamental connection btween regexes and parsing. That is to say, regexes correspond to a particular subset of grammars describe only regular languages. An example of a grammar which describes a regular language:
S -> A | B
B -> b
A -> A a | c
Although there is A -> A ... recursion, this is still regular, and corresponds to the regex ac*|b, which is just a more compact way to denote the same language. The grammar lets us notate languages that aren't regular and for which we can't write a regex, but as we have seen, we can extend the regex notation and semantics to express some of these things. Regular expressions aren't separate from grammars. The two aren't counterparts, but rather one is a special case or subset of the other.
Parser generators like Yacc, Bison, and derivatives are what you're after. They aren't as widespread as regular expressions because they generate actual C code. There are translations like Jison for example which implement the Yacc/Bison syntax using javascript. I know there are similar tools for other languages.
I get the impression Parsing expression grammar systems are up and coming though.

Is there a regular language to represent regular expressions?

Specifically, I noticed that the language of regular expressions itself isn't regular. So, I can't use a regular expression to parse a given regular expression. I need to use a parser since the language of the regular expression itself is context free.
Is there any way regular expressions can be represented in a way that the resulting string can be parsed using a regular expression?
Note: My question isn't about whether there is a regexp to match the current syntax of regexes, but whether there exists a "representation" for regular expressions as we know it today (maybe not a neat as what we know them as today) that can be parsed using regular expressions. Also, please could someone remove the dup since it isn't a dup. I'm asking something completely different. I already know that the current language of regular expressions isn't regular (it is how I started my original question).
Depending on what you mean by "represent", the answer is "yes" or "no":
If you want a language that (homomorphically) maps 1:1 to the usual basic regular expression language, the answer is no, because a regular language cannot be isomorphic to a non-regular language, and the standard regular expression language is non-regular. This is because the syntax requires matching opening and closing parentheses of arbitrary depth.
If "represent" only means another method of specifying regular languages, the answer is yes, and right now I can think of at least three ways to achieve this:
The "dumbest" and easiest way is to define some surjective mapping f : ℕ -> RegEx from the natural numbers onto the set of all valid standard regular expressions. You can define the natural numbers using the regular expression 0|1[01]*, and the regular language denoted by a (string representing the) natural number n is the regular language denoted by f(n).
Of course, the meaning attached to a natural number would not be obvious to a human reader at all, so this "regular expression language" would be utterly useless.
As parentheses are the only non-regular part in simple regular expressions, the easiest human-interpretable method would be to extend the standard simple regular expression syntax to allow dangling parentheses and defining semantics for dangling parentheses.
The obvious choice would be to ignore non-matching opening parentheses and interpreting non-matching closing parentheses as matching the beginning of the regex. This essentially amounts to implicitly inserting as many opening parentheses at the beginning and as many closing parentheses at the end of the regex as necessary. Additionally, (* would have to be interpreted as repetition of the empty string. If I didn't miss anything, this definition should turn any string into a "regular expression" with a specified meaning, so .* defines this "regular expression language".
This variant even has the same abstract syntax as standard regular expressions.
Another variant would be to specify the NFA that recognizes the language directly using a regular language, e.g.: ([a-z]+,([^,]|\\,|\\\\)+,[a-z]+\$?;)*.
The idea is that [a-z]+ is used as a label for states, and the expression is a list of transition triples (s, c, t) from source state s to target state t consuming character c, and a $ indicating accepting transitions (cf. note below). In c, backslashes are used to escape commas or backslashes - I assumed that you use the same alphabet for standard regular expressions, but of course you can replace the middle component with any other regular language of symbols denotating characters of any alphabet you wish.
The first source state mentioned is the (single) initial state. An empty expression defines the empty language.
Above, I wrote "accepting transition", not "accepting state" because that would make the regex above a bit more complex. You can interpret a triple containing a $ as two transitions, namely one transition consuming c from s to a new, unique state, and an ε-transition from that state to t. This should allow any NFA to be represented, by replacing each transition to an accepting state with a $ triple and each transition to a non-accepting state with a non-$ triple.
One note that might make the "yes" part look more intuitive: Assembly languages are regular, and those are even Turing-complete, so it would be unexpected if it wasn't possible to specify "mere" regular languages using a regular language.
The answer is probably NO.
As you have pointed out, set of all possible regular expressions itself is not a regular set. Any TRUE regular expression (not those extended) can be converted into finite automata (FA). If regular expression can be represented in a form that can be parsed by itself, then FA can be parsed by regular expression as well.
But that's not possible as far as I know. RE itself can be reduced into three basic operation(According to the Dragon Book):
concatenation: e.g. ab
alternation: e.g. a|b
kleen closure: e.g. a*
The kleen closure can match infinite number of characters, but it cannot know how many characters to match.
Just think such case: you want to match 3 consecutive as. Then the corresponding regular expression is /aaa/. But what if you want match 4, 5, 6... as? Parser with only one RE cannot know the exact number of as. So it fails to give the right matching to arbitrary expressions. However, the RE parser has to match infinite different forms of REs. According to your expression, a regular expression cannot match all the possibilities.
Well, the only difference of a RE parser is that it does not need a tokenizer.(probably that's why RE is used in lexical analysis) Every character in RE is a token (excluding those escape charcters). But to parse RE, whatever it is converted,one has to face up with NFA/DFA/TREE... all equivalent structures that cannot be parsed by RE itself.

Aren't modern regular expression dialects regular?

I've seen a few comments here that mention that modern regular expressions go beyond what can be represented in a regular language. How is this so?
What features of modern regular expressions are not regular? Examples would be helpful.
The first thing that comes to mind is backreferences:
(\w*)\s\1
(matches a group of word characters, followed by a space character and then the same group previously matched) eg: hello hello matches, hello world doesn't.
This construct is not regular (ie: can't be generated by a regular grammar).
Another feature supported by Perl Compatible RegExp (PCRE) that is not regular are recursive patterns:
\((a*|(?R))*\)
This can be used to match any combination of balanced parentheses and "a"s (from wikipedia)
A couple of examples:
Regular expressions support grouping. E.g. in Ruby: /my (group)/.match("my group")[1] will output "group". storing something in a group requires an external storage, which a finite automaton does not have.
Many languages, e.g. C#, support captures, i.e. that each match will be captured on a stack - for example the pattern (?<MYGROUP>.)* could perform multiple captures of "." in the same group.
Grouping are used for backreferencing as pointed out by the user NullUserException above. Backreferencing requires one or more external stacks with the power of a push-down-automaton (you have to be able to push something on a stack and peek or pop it afterwards.
Some engines have the possibility of seperately pushing and popping external stacks and checking whether the stack is empty. In .NET, actually (?<MYGROUP>test) pushes a stack, while (?<-MYGROUP>) pops a stack.
Some engines like the .NET engine have a balanced grouping concept - where an external stack can be both pushed and popped at the same time. Balanced grouping syntax is (?<FIRSTGROUP-LASTGROUP>) which pops the LASTGROUP and pushes the capture since the LASTGROUP index on the FIRSTGROUP stack. This can actually be used to match infinitely nested constructions which is definitely beyond the power of a finite automaton.
Probably other good examples exist :-) If you are further interessted in some of the implementation details of external stacks in combination with Regex's and balanced grouping and thus higher order automata than finite automata, I once wrote two short articles on this (http://www.codeproject.com/KB/recipes/Nested_RegEx_explained.aspx and http://www.codeproject.com/KB/recipes/RegEx_Balanced_Grouping.aspx).
Anyway - finitieness or not - I blieve that the power that this extra stuff brings to the regular languages is great :-)
Br. Morten
A deterministic or nondeterministic finite automaton recognizes just the regular languages, which are described by regular expressions. The definition of a regular expression is simple. Let S be an alphabet. Then the empty set, the empty string, and every element of S are regular expressions (over S). Let u and v be regular expressions. Then the union (u | v), concatenation (uv), and closure (u*) of u and v are regular expressions over S. This definition is easily extended to the regular languages. No other expression is a regular expression. As pointed out, some back-references are an example. The Wikipedia pages on regular languages and expressions are good references.
In essence, certain "regular expressions" are not regular because no automaton of a particular type can be constructed to recognize them. For example, the the language
{ a^ i b^ i : i <= 0 }
is not regular. This is because the accepting automaton would require infinitely many states, but an automaton accepting regular languages must have a finite number of states.

Does lookaround affect which languages can be matched by regular expressions?

There are some features in modern regex engines which allow you to match languages that couldn't be matched without that feature. For example the following regex using back references matches the language of all strings that consist of a word that repeats itself: (.+)\1. This language is not regular and can't be matched by a regex that does not use back references.
Does lookaround also affect which languages can be matched by a regular expression? I.e. are there any languages that can be matched using lookaround that couldn't be matched otherwise? If so, is this true for all flavors of lookaround (negative or positive lookahead or lookbehind) or just for some of them?
The answer to the question you ask, which is whether a larger class of languages than the regular languages can be recognised with regular expressions augmented by lookaround, is no.
A proof is relatively straightforward, but an algorithm to translate a regular expression containing lookarounds into one without is messy.
First: note that you can always negate a regular expression (over a finite alphabet). Given a finite state automaton that recognises the language generated by the expression, you can simply exchange all the accepting states for non-accepting states to get an FSA that recognises exactly the negation of that language, for which there are a family of equivalent regular expressions.
Second: because regular languages (and hence regular expressions) are closed under negation they are also closed under intersection since A intersect B = neg ( neg(A) union neg(B)) by de Morgan's laws. In other words given two regular expressions, you can find another regular expression that matches both.
This allows you to simulate lookaround expressions. For example u(?=v)w matches only expressions that will match uv and uw.
For negative lookahead you need the regular expression equivalent of the set theoretic A\B, which is just A intersect (neg B) or equivalently neg (neg(A) union B). Thus for any regular expressions r and s you can find a regular expression r-s which matches those expressions that match r which do not match s. In negative lookahead terms: u(?!v)w matches only those expressions which match uw - uv.
There are two reasons why lookaround is useful.
First, because the negation of a regular expression can result in something much less tidy. For example q(?!u)=q($|[^u]).
Second, regular expressions do more than match expressions, they also consume characters from a string - or at least that's how we like to think about them. For example in python I care about the .start() and .end(), thus of course:
>>> re.search('q($|[^u])', 'Iraq!').end()
5
>>> re.search('q(?!u)', 'Iraq!').end()
4
Third, and I think this is a pretty important reason, negation of regular expressions does not lift nicely over concatenation. neg(a)neg(b) is not the same thing as neg(ab), which means that you cannot translate a lookaround out of the context in which you find it - you have to process the whole string. I guess that makes it unpleasant for people to work with and breaks people's intuitions about regular expressions.
I hope I have answered your theoretical question (its late at night, so forgive me if I am unclear). I agree with a commentator who said that this does have practical applications. I met very much the same problem when trying to scrape some very complicated web pages.
EDIT
My apologies for not being clearer: I do not believe you can give a proof of regularity of regular expressions + lookarounds by structural induction, my u(?!v)w example was meant to be just that, an example, and an easy one at that. The reason a structural induction won't work is because lookarounds behave in a non-compositional way - the point I was trying to make about negations above. I suspect any direct formal proof is going to have lots of messy details. I have tried to think of an easy way to show it but cannot come up with one off the top of my head.
To illustrate using Josh's first example of ^([^a]|(?=..b))*$ this is equivalent to a 7 state DFSA with all states accepting:
A - (a) -> B - (a) -> C --- (a) --------> D
Λ | \ |
| (not a) \ (b)
| | \ |
| v \ v
(b) E - (a) -> F \-(not(a)--> G
| <- (b) - / |
| | |
| (not a) |
| | |
| v |
\--------- H <-------------------(b)-----/
The regular expression for state A alone looks like:
^(a([^a](ab)*[^a]|a(ab|[^a])*b)b)*$
In other words any regular expression you are going to get by eliminating lookarounds will in general be much longer and much messier.
To respond to Josh's comment - yes I do think the most direct way to prove the equivalence is via the FSA. What makes this messier is that the usual way to construct an FSA is via a non-deterministic machine - its much easier to express u|v as simply the machine constructed from machines for u and v with an epsilon transition to the two of them. Of course this is equivalent to a deterministic machine, but at the risk of exponential blow-up of states. Whereas negation is much easier to do via a deterministic machine.
The general proof will involve taking the cartesian product of two machines and selecting those states you wish to retain at each point you want to insert a lookaround. The example above illustrates what I mean to some extent.
My apologies for not supplying a construction.
FURTHER EDIT:
I have found a blog post which describes an algorithm for generating a DFA out of a regular expression augmented with lookarounds. Its neat because the author extends the idea of an NFA-e with "tagged epsilon transitions" in the obvious way, and then explains how to convert such an automaton into a DFA.
I thought something like that would be a way to do it, but I'm pleased that someone has written it up. It was beyond me to come up with something so neat.
As the other answers claim, lookarounds don't add any extra power to regular expressions.
I think we can show this using the following:
One Pebble 2-NFA (see the Introduction section of the paper which refers to it).
The 1-pebble 2NFA does not deal with nested lookaheads, but, we can use a variant of multi-pebble 2NFAs (see section below).
Introduction
A 2-NFA is a non deterministic finite automaton which has the ability to move either left or right on it's input.
A one pebble machine is where the machine can place a pebble on the input tape (i.e. mark a specific input symbol with a pebble) and do possibly different transitions based on whether there is a pebble at the current input position or not.
It is known the One Pebble 2-NFA has the same power as a regular DFA.
Non-nested Lookaheads
The basic idea is as follows:
The 2NFA allows us to backtrack (or 'front track') by moving forward or backward in the input tape. So for a lookahead we can do the match for the lookahead regular expression and then backtrack what we have consumed, in matching the lookahead expression. In order to know exactly when to stop backtracking, we use the pebble! We drop the pebble before we enter the dfa for the lookahead to mark the spot where the backtracking needs to stop.
Thus at the end of running our string through the pebble 2NFA, we know whether we matched the lookahead expression or not and the input left (i.e. what is left to be consumed) is exactly what is required to match the remaining.
So for a lookahead of the form u(?=v)w
We have the DFAs for u, v and w.
From the accepting state (yes, we can assume there is only one) of DFA for u, we make an e-transition to the start state of v, marking the input with a pebble.
From an accepting state for v, we e-transtion to a state which keeps moving the input left, till it finds a pebble, and then transitions to start state of w.
From a rejecting state of v, we e-transition to a state which keeps moving left until it finds the pebble, and transtions to the accepting state of u (i.e where we left off).
The proof used for regular NFAs to show r1 | r2, or r* etc, carry over for these one pebble 2nfas. See http://www.coli.uni-saarland.de/projects/milca/courses/coal/html/node41.html#regularlanguages.sec.regexptofsa for more info on how the component machines are put together to give the bigger machine for the r* expression etc.
The reason why the above proofs for r* etc work is that the backtracking ensures that the input pointer is always at the right spot, when we enter the component nfas for repetition. Also, if a pebble is in use, then it is being processed by one of the lookahead component machines. Since there are no transitions from lookahead machine to lookahead machine without completely backtracking and getting back the pebble, a one pebble machine is all that is needed.
For eg consider ([^a] | a(?=...b))*
and the string abbb.
We have abbb which goes through the peb2nfa for a(?=...b), at the end of which we are at the state: (bbb, matched) (i.e in input bbb is remaining, and it has matched 'a' followed by '..b'). Now because of the *, we go back to the beginning (see the construction in the link above), and enter the dfa for [^a]. Match b, go back to beginning, enter [^a] again two times, and then accept.
Dealing with Nested Lookaheads
To handle nested lookaheads we can use a restricted version of k-pebble 2NFA as defined here: Complexity Results for Two-Way and Multi-Pebble Automata and their Logics (see Definition 4.1 and Theorem 4.2).
In general, 2 pebble automata can accept non-regular sets, but with the following restrictions, k-pebble automata can be shown to be regular (Theorem 4.2 in above paper).
If the pebbles are P_1, P_2, ..., P_K
P_{i+1} may not be placed unless P_i is already on the tape and P_{i} may not be picked up unless P_{i+1} is not on the tape. Basically the pebbles need to be used in a LIFO fashion.
Between the time P_{i+1} is placed and the time that either P_{i} is picked up or P_{i+2} is placed, the automaton can traverse only the subword located between the current location of P_{i} and the end of the input word that lies in the direction of P_{i+1}. Moreover, in this sub-word, the automaton can act only as a 1-pebble automaton with Pebble P_{i+1}. In particular it is not allowed to lift up, place or even sense the presence of another pebble.
So if v is a nested lookahead expression of depth k, then (?=v) is a nested lookahead expression of depth k+1. When we enter a lookahead machine within, we know exactly how many pebbles have to have been placed so far and so can exactly determine which pebble to place and when we exit that machine, we know which pebble to lift. All machines at depth t are entered by placing pebble t and exited (i.e. we return to processing of a depth t-1 machine) by removing pebble t. Any run of the complete machine looks like a recursive dfs call of a tree and the above two restrictions of the multi-pebble machine can be catered to.
Now when you combine expressions, for rr1, since you concat, the pebble numbers of r1 must be incremented by the depth of r. For r* and r|r1 the pebble numbering remains the same.
Thus any expression with lookaheads can be converted to an equivalent multi-pebble machine with the above restrictions in pebble placement and so is regular.
Conclusion
This basically addresses the drawback in Francis's original proof: being able to prevent the lookahead expressions from consuming anything which are required for future matches.
Since Lookbehinds are just finite string (not really regexs) we can deal with them first, and then deal with the lookaheads.
Sorry for the incomplete writeup, but a complete proof would involve drawing a lot of figures.
It looks right to me, but I will be glad to know of any mistakes (which I seem to be fond of :-)).
I agree with the other posts that lookaround is regular (meaning that it does not add any fundamental capability to regular expressions), but I have an argument for it that is simpler IMO than the other ones I have seen.
I will show that lookaround is regular by providing a DFA construction. A language is regular if and only if it has a DFA that recognizes it. Note that Perl doesn't actually use DFAs internally (see this paper for details: http://swtch.com/~rsc/regexp/regexp1.html) but we construct a DFA for purposes of the proof.
The traditional way of constructing a DFA for a regular expression is to first build an NFA using Thompson's Algorithm. Given two regular expressions fragments r1 and r2, Thompson's Algorithm provides constructions for concatenation (r1r2), alternation (r1|r2), and repetition (r1*) of regular expressions. This allows you to build a NFA bit by bit that recognizes the original regular expression. See the paper above for more details.
To show that positive and negative lookahead are regular, I will provide a construction for concatenation of a regular expression u with positive or negative lookahead: (?=v) or (?!v). Only concatenation requires special treatment; the usual alternation and repetition constructions work fine.
The construction is for both u(?=v) and u(?!v) is:
In other words, connect every final state of the existing NFA for u to both an accept state and to an NFA for v, but modified as follows. The function f(v) is defined as:
Let aa(v) be a function on an NFA v that changes every accept state into an "anti-accept state". An anti-accept state is defined to be a state that causes the match to fail if any path through the NFA ends in this state for a given string s, even if a different path through v for s ends in an accept state.
Let loop(v) be a function on an NFA v that adds a self-transition on any accept state. In other words, once a path leads to an accept state, that path can stay in the accept state forever no matter what input follows.
For negative lookahead, f(v) = aa(loop(v)).
For positive lookahead, f(v) = aa(neg(v)).
To provide an intuitive example for why this works, I will use the regex (b|a(?:.b))+, which is a slightly simplified version of the regex I proposed in the comments of Francis's proof. If we use my construction along with the traditional Thompson constructions, we end up with:
The es are epsilon transitions (transitions that can be taken without consuming any input) and the anti-accept states are labeled with an X. In the left half of the graph you see the representation of (a|b)+: any a or b puts the graph in an accept state, but also allows a transition back to the begin state so we can do it again. But note that every time we match an a we also enter the right half of the graph, where we are in anti-accept states until we match "any" followed by a b.
This is not a traditional NFA because traditional NFAs don't have anti-accept states. However we can use the traditional NFA->DFA algorithm to convert this into a traditional DFA. The algorithm works like usual, where we simulate multiple runs of the NFA by making our DFA states correspond to subsets of the NFA states we could possibly be in. The one twist is that we slightly augment the rule for deciding if a DFA state is an accept (final) state or not. In the traditional algorithm a DFA state is an accept state if any of the NFA states was an accept state. We modify this to say that a DFA state is an accept state if and only if:
= 1 NFA states is an accept state, and
0 NFA states are anti-accept states.
This algorithm will give us a DFA that recognizes the regular expression with lookahead. Ergo, lookahead is regular. Note that lookbehind requires a separate proof.
I have a feeling that there are two distinct questions being asked here:
Are Regex engines that encorporate "lookaround" more
powerful than Regex engines that don't?
Does "lookaround"
empower a Regex engine with the ability to parse languages that are
more complex than those generated from a Chomsky Type 3 - Regular grammar?
The answer to the first question in a practical sense is yes. Lookaround will give a Regex engine that
uses this feature fundamentally more power than one that doesn't. This is because
it provides a richer set of "anchors" for the matching process.
Lookaround lets you define an entire Regex as a possible anchor point (zero width assertion). You can
get a pretty good overview of the power of this feature here.
Lookaround, although powerful, does not lift the Regex engine beyond the theoretical
limits placed on it by a Type 3 Grammar. For example, you will never be able to reliably
parse a language based on a Context Free - Type 2 Grammar using a Regex engine
equipped with lookaround. Regex engines are limited to the power of a Finite State Automation
and this fundamentally restricts the expressiveness of any language they can parse to the level of a Type 3 Grammar. No matter
how many "tricks" are added to your Regex engine, languages generated via a Context Free Grammar
will always remain beyond its capabilities. Parsing Context Free - Type 2 grammar requires pushdown automation to "remember" where it is in
a recursive language construct. Anything that requires a recursive evaluation of the grammar rules cannot be parsed using
Regex engines.
To summarize: Lookaround provides some practical benefits to Regex engines but does not "alter the game" on a
theoretical level.
EDIT
Is there some grammar with a complexity somewhere between Type 3 (Regular) and Type 2 (Context Free)?
I believe the answer is no. The reason is because there is no theoretical limit
placed on the size of the NFA/DFA needed to describe a Regular language. It may become arbitrarily large
and therefore impractical to use (or specify). This is where dodges such as "lookaround" are useful. They
provide a short-hand mechanism to specify what would otherwise lead to very large/complex NFA/DFA
specifications. They do not increase the expressiveness of
Regular languages, they just make specifying them more practical. Once you get this point, it becomes
clear that there are a lot of "features" that could be added to Regex engines to make them more
useful in a practical sense - but nothing will make them capable of going beyond the
limits of a Regular language.
The basic difference between a Regular and a Context Free language is that a Regular language
does not contain recursive elements. In order to evaluate a recursive language you need a
Push Down Automation
to "remember" where you are in the recursion. An NFA/DFA does not stack state information so cannot
handle the recursion. So given a non-recursive language definition there will be some NFA/DFA (but
not necessarily a practical Regex expression) to describe it.

DFA based regular expression matching - how to get all matches?

I have a given DFA that represent a regular expression.
I want to match the DFA against an input stream and get all possible matches back, not only the leastmost-longest match.
For example:
regex: a*ba|baa
input: aaaaabaaababbabbbaa
result:
aaaaaba
aaba
ba
baa
Assumptions
Based on your question and later comments you want a general method for splitting a sentence into non-overlapping, matching substrings, with non-matching parts of the sentence discarded. You also seem to want optimal run-time performance. Also I assume you have an existing algorithm to transform a regular expression into DFA form already. I further assume that you are doing this by the usual method of first constructing an NFA and converting it by subset construction to DFA, since I'm not aware of any other way of accomplishing this.
Before you go chasing after shadows, make sure your trying to apply the right tool for the job. Any discussion of regular expressions is almost always muddied by the fact that folks use regular expressions for a lot more things than they are really optimal for. If you want to receive the benefits of regular expressions, be sure you're using a regular expression, and not something broader. If what you want to do can't be somewhat coded into a regular expression itself, then you can't benefit from the advantages of regular expression algorithms (fully)
An obvious example is that no amount of cleverness will allow a FSM, or any algorithm, to predict the future. For instance, an expression like (a*b)|(a), when matched against the string aaa... where the ellipsis is the portion of the expression not yet scanned because the user has not typed them yet, cannot give you every possible right subgroup.
For a more detailed discussion of Regular expression implementations, and specifically Thompson NFA's please check this link, which describes a simple C implementation with some clever optimizations.
Limitations of Regular Languages
The O(n) and Space(O(1)) guarantees of regular expression algorithms is a fairly narrow claim. Specifically, a regular language is the set of all languages that can be recognized in constant space. This distinction is important. Any kind of enhancement to the algorithm that does something more sophisticated than accepting or rejecting a sentence is likely to operate on a larger set of languages than regular. On top of that, if you can show that some enhancement requires greater than constant space to implement, then you are also outside of the performance guarantee. That being said, we can still do an awful lot if we are very careful to keep our algorithm within these narrow constraints.
Obviously that eliminates anything we might want to do with recursive backtracking. A stack does not have constant space. Even maintaining pointers into the sentence would be verboten, since we don't know how long the sentence might be. A long enough sentence would overflow any integer pointer. We can't create new states for the automaton as we go to get around this. All possible states (and a few impossible ones) must be predictable before exposing the recognizer to any input, and that quantity must be bounded by some constant, which may vary for the specific language we want to match, but by no other variable.
This still allows some room for adding additonal behavior. The usual way of getting more mileage is to add some extra annotations for where certain events in processing occur, such as when a subexpression started or stopped matching. Since we are only allowed to have constant space processing, that limits the number of subexpression matches we can process. This usually means the latest instance of that subexpression. This is why, when you ask for the subgrouped matched by (a|)*, you always get an empty string, because any sequence of a's is implicitly followed by infinitely many empty strings.
The other common enhancement is to do some clever thing between states. For example, in perl regex, \b matches the empty string, but only if the previous character is a word character and the next is not, or visa versa. Many simple assertions fit this, including the common line anchor operators, ^ and $. Lookahead and lookbehind assertions are also possible, but much more difficult.
When discussing the differences between various regular language recognizers, it's worth clarifying if we're talking about match recognition or search recognition, the former being an accept only if the entire sentence is in the language, and the latter accepts if any substring in the sentence is in the language. These are equivalent in the sense that if some expression E is accepted by the search method, then .*(E).* is accepted in the match method.
This is important because we might want to make it clear whether an expression like a*b|a accepts aa or not. In the search method, it does. Either token will match the right side of the disjunction. It doesn't match, though, because you could never get that sentence by stepping through the expression and generating tokens from the transitions, at least in a single pass. For this reason, i'm only going to talk about match semantics. Obviously if you want search semantics, you can modify the expression with .*'s
Note: A language defined by expression E|.* is not really a very manageable language, regardless of the sublanguage of E because it matches all possible sentences. This represents a real challenge for regular expression recognizers because they are really only suited to recognizing a language or else confirming that a sentence is not in that same language, rather than doing any more specific work.
Implementation of Regular Language Recognizers
There are generally three ways to process a regular expression. All three start the same, by transforming the expression into an NFA. This process produces one or two states for each production rule in the original expression. The rules are extemely simple. Here's some crude ascii art: note that a is any single literal character in the language's alphabet, and E1 and E2 are any regular expression. Epsilon(ε) is a state with inputs and outputs, but ignores the stream of characters, and doesn't consume any input either.
a ::= > a ->
E1 E2 ::= >- E1 ->- E2 ->
/---->
E1* ::= > --ε <-\
\ /
E1
/-E1 ->
E1|E2 ::= > ε
\-E2 ->
And that's it! Common uses such as E+, E?, [abc] are equivalent to EE*, (E|), (a|b|c) respectively. Also note that we add for each production rule a very small number of new states. In fact each rule adds zero or one state (in this presentation). characters, quantifiers and dysjunction all add just one state, and the concatenation doesn't add any. Everything else is done by updating the fragments' end pointers to start pointers of other states or fragments.
The epsilon transition states are important, because they are ambiguous. When encountered, is the machine supposed to change state to once following state or another? should it change state at all or stay put? That's the reason why these automatons are called nondeterministic. The solution is to have the automaton transition to the right state, whichever allows it to match the best. Thus the tricky part is to figure out how to do that.
There are fundamentally two ways of doing this. The first way is to try each one. Follow the first choice, and if that doesn't work, try the next. This is recursive backtracking, appears in a few (and notable) implementations. For well crafted regular expressions, this implementation does very little extra work. If the expression is a bit more convoluted, recursive backtracking is very, very bad, O(2^n).
The other way of doing this is to instead try both options in parallel. At each epsilon transition, add to the set of current states both of the states the epsilon transition suggests. Since you are using a set, you can have the same state come up more than once, but you only need to track it once, either you are in that state or not. If you get to the point that there's no option for a particular state to follow, just ignore it, that path didn't match. If there are no more states, then the entire expression didn't match. as soon as any state reaches the final state, you are done.
Just from that explanation, the amount of work we have to do has gone up a little bit. We've gone from having to keep track of a single state to several. At each iteration, we may have to update on the order of m state pointers, including things like checking for duplicates. Also the amount of storage we needed has gone up, since now it's no longer a single pointer to one possible state in the NFA, but a whole set of them.
However, this isn't anywhere close to as bad as it sounds. First off, the number of states is bounded by the number of productions in the original regular expression. From now on we'll call this value m to distinguish it from the number of symbols in the input, which will be n. If two state pointers end up transitioning to the same new state, you can discard one of them, because no matter what else happens, they will both follow the same path from there on out. This means the number of state pointers you need is bounded by the number of states, so that to is m.
This is a bigger win in the worst case scenario when compared to backtracking. After each character is consumed from the input, you will create, rename, or destroy at most m state pointers. There is no way to craft a regular expression which will cause you to execute more than that many instructions (times some constant factor depending on your exact implementation), or will cause you to allocate more space on the stack or heap.
This NFA, simultaneously in some subset of its m states, may be considered some other state machine who's state represents the set of states the NFA it models could be in. each state of that FSM represents one element from the power set of the states of the NFA. This is exactly the DFA implementation used for matching regular expressions.
Using this alternate representation has an advantage that instead of updating m state pointers, you only have to update one. It also has a downside, since it models the powerset of m states, it actually has up to 2m states. That is an upper limit, because you don't model states that cannot happen, for instance the expression a|b has two possible states after reading the first character, either the one for having seen an a, or the one for having seen a b. No matter what input you give it, it cannot be in both of those states at the same time, so that state-set does not appear in the DFA. In fact, because you are eliminating the redundancy of epsilon transitions, many simple DFA's actually get SMALLER than the NFA they represent, but there is simply no way to guarantee that.
To keep the explosion of states from growing too large, a solution used in a few versions of that algorithm is to only generate the DFA states you actually need, and if you get too many, discard ones you haven't used recently. You can always generate them again.
From Theory to Practice
Many practical uses of regular expressions involve tracking the position of the input. This is technically cheating, since the input could be arbitrarily long. Even if you used a 64 bit pointer, the input could possibly be 264+1 symbols long, and you would fail. Your position pointers have to grow with the length of the input, and now your algorithm now requires more than constant space to execute. In practice this isn't relevant, because if your regular expression did end up working its way through that much input, you probably won't notice that it would fail because you'd terminate it long before then.
Of course, we want to do more than just accept or reject inputs as a whole. The most useful variation on this is to extract submatches, to discover which portion of an input was matched by a certain section of the original expression. The simple way to achieve this is to add an epsilon transition for each of the opening and closing braces in the expression. When the FSM simulator encounters one of these states, it annotates the state pointer with information about where in the input it was at the time it encountered that particular transition. If the same pointer returns to that transition a second time, the old annotation is discarded and replaced with a new annotation for the new input position. If two states pointers with disagreeing annotations collapse to the same state, the annotation of a later input position wins again.
If you are sticking to Thompson NFA or DFA implementations, then there's not really any notion of greedy or non-greedy matching. A backtracking algorithm needs to be given a hint about whether it should start by trying to match as much as it can and recursively trying less, or trying as little as it can and recursively trying more, when it fails it first attempt. The Thompson NFA method tries all possible quantities simultaneously. On the other hand, you might still wish to use some greedy/nongreedy hinting. This information would be used to determine if newer or older submatch annotations should be preferred, in order to capture just the right portion of the input.
Another kind of practical enhancement is assertions, productions which do not consume input, but match or reject based on some aspect of the input position. For instance in perl regex, a \b indicates that the input must contain a word boundary at that position, such that the symbol just matched must be a word character, but the next character must not be, or visa versa. Again, we manage this by adding an epsilon transition with special instructions to the simulator. If the assertion passes, then the state pointer continues, otherwise it is discarded.
Lookahead and lookbehind assertions can be achieved with a bit more work. A typical lookbehind assertion r0(?<=r1)r2 is transformed into two separate expressions, .*r1 and r0εr2. Both expressions are applied to the input. Note that we added a .* to the assertion expression, because we don't actually care where it starts. When the simulator encounters the epsilon in the second generated fragment, it checks up on the state of the first fragment. If that fragment is in a state where it could accept right there, the assertion passes with the state pointer flowing into r2, but otherwise, it fails, and both fragments continue, with the second discarding the state pointer at the epsilon transition.
Lookahead also works by using an extra regex fragment for the assertion, but is a little more complex, because when we reach the point in the input where the assertion must succeed, none of the corresponding characters have been encountered (in the lookbehind case, they have all been encountered). Instead, when the simulator reaches the assertion, it starts a pointer in the start state of the assertion subexpression and annotates the state pointer in the main part of the simulation so that it knows it is dependent on the subexpression pointer. At each step, the simulation must check to see that the state pointer it depends upon is still matching. If it doesn't find one, then it fails wherever it happens to be. You don't have to keep any more copies of the assertion subexpressions state pointers than you do for the main part, if two state pointers in the assertion land on the same state, then the state pointers each of them depend upon will share the same fate, and can be reannotated to point to the single pointer you keep.
While were adding special instructions to epsilon transitions, it's not a terrible idea to suggest an instruction to make the simulator pause once in a while to let the user see what's going on. Whenever the simulator encounters such a transition, it will wrap up its current state in some kind of package that can be returned to the caller, inspected or altered, and then resumed where it left off. This could be used to match input interactively, so if the user types only a partial match, the simulator can ask for more input, but if the user types something invalid, the simulator is empty, and can complain to the user. Another possibility is to yield every time a subexpression is matched, allowing you to peek at every sub match in the input. This couldn't be used to exclude some submatches, though. For instance, if you tried to match ((a)*b) against aaa, you could see three submatches for the a's, even though the whole expression ultimately fails because there is no b, and no submatch for the corresponding b's
Finally, there might be a way to modify this to work with backreferences. Even if it's elegent, it's sure to be inefficient, specifically, regular expressions plus backreferences are in NP-Complete, so I won't even try to think of a way to do this, because we are only interested (here) in (asymptotically) efficient possibilities.