Printed from Urbana-Champaign IMC : http://www.ucimc.org/
UCIMC Independent Media 
Center
Media Centers

[topics]
biotech

[regions]
united states

oceania

germany

[projects]
video
satellite tv
radio
print

[process]
volunteer
tech
process & imc docs
mailing lists
indymedia faq
fbi/legal updates
discussion

west asia
palestine
israel
beirut

united states
worcester
western mass
virginia beach
vermont
utah
urbana-champaign
tennessee
tampa bay
tallahassee-red hills
seattle
santa cruz, ca
santa barbara
san francisco bay area
san francisco
san diego
saint louis
rogue valley
rochester
richmond
portland
pittsburgh
philadelphia
omaha
oklahoma
nyc
north texas
north carolina
new orleans
new mexico
new jersey
new hampshire
minneapolis/st. paul
milwaukee
michigan
miami
maine
madison
la
kansas city
ithaca
idaho
hudson mohawk
houston
hawaii
hampton roads, va
dc
danbury, ct
columbus
colorado
cleveland
chicago
charlottesville
buffalo
boston
binghamton
big muddy
baltimore
austin
atlanta
arkansas
arizona

south asia
mumbai
india

oceania
sydney
perth
melbourne
manila
jakarta
darwin
brisbane
aotearoa
adelaide

latin america
valparaiso
uruguay
tijuana
santiago
rosario
qollasuyu
puerto rico
peru
mexico
ecuador
colombia
chile sur
chile
chiapas
brasil
bolivia
argentina

europe
west vlaanderen
valencia
united kingdom
ukraine
toulouse
thessaloniki
switzerland
sverige
scotland
russia
romania
portugal
poland
paris/ãŽle-de-france
oost-vlaanderen
norway
nice
netherlands
nantes
marseille
malta
madrid
lille
liege
la plana
italy
istanbul
ireland
hungary
grenoble
galiza
euskal herria
estrecho / madiaq
cyprus
croatia
bulgaria
bristol
belgrade
belgium
belarus
barcelona
austria
athens
armenia
antwerpen
andorra
alacant

east asia
qc
japan
burma

canada
winnipeg
windsor
victoria
vancouver
thunder bay
quebec
ottawa
ontario
montreal
maritimes
london, ontario
hamilton

africa
south africa
nigeria
canarias
ambazonia

www.indymedia.org

This site
made manifest by
dadaIMC software
&
the friendly folks of
AcornActiveMedia.com

Comment on this article | Email this Article
News :: Miscellaneous
Computational language Current rating: 0
07 Jun 2002
Modified: 09:52:05 PM
Language is our legacy. It is the main evolutionary contribution of humans, and perhaps the most interesting trait that has emerged in the past 500 million years. Understanding how darwinian evolution gives rise to human language requires the integration of formal language theory, learning theory and evolutionary dynamics. Formal language theory provides a mathematical description of language and grammar. Learning theory formalizes the task of language acquisition—it can be shown that no procedure can learn an unrestricted set of languages. Universal grammar specifies the restricted set of languages learnable by the human brain. Evolutionary dynamics can be formulated to describe the cultural evolution of language and the biological evolution of universal grammar.







Biology uses generative systems. Genomes consist of an alphabet of four nucleotides, which, together with certain rules for how to produce proteins and organize cells, generates an unlimited variety of living organisms. For more than 3 billion years, evolution of life on Earth was restricted to using this generative system. Only very recently another generative system emerged, which led to a new mode of evolution. This other system is human language. It enables us to transfer unlimited non-genetic information among individuals, and it gives rise to cultural evolution.

Currently there are many efforts to bring linguistic inquiry into contact with several areas of biology including evolution1-11, genetics12-14, neurobiology15, 16 and animal behaviour17-20. The aim of this Review is to formulate a synthesis of formal language theory21, 22, learning theory23-28 and evolutionary dynamics in a manner that is useful for people from various disciplines. We will address the following questions: What is language? What is grammar? What is learning? How does a child learn language? What is the difference between learning language and learning other generative systems? In what sense is there a logical necessity for genetically determined components of human language, such as 'universal grammar'? Finally, we will discuss how formal language theory and learning theory can be extended to study language as a biological phenomenon, as a product of evolution.

Formal language theory
Language is a mode of communication, a crucial part of human behaviour and a cultural object defining our social identity. There is also a fundamental aspect of human language that makes it amenable to formal analysis: linguistic structures consist of smaller units that are grouped together according to certain rules.

The combinatorial sequencing of small units into bigger structures occurs at several different levels. Phonemes form syllables and words. Words form phrases and sentences. The rules for such groupings are not arbitrary. Any native English speaker recognizes that the sentence 'He ran from there with his money' obeys the rules of English, while 'He his money with there from ran' does not. In Bengali the reverse is true.

Individual languages have specific rules. Certain word orders are admissible in one language but not in another. In some languages, word order is relatively free but case marking is pronounced. There are always specific rules that generate valid or meaningful linguistic structures. Much of modern linguistic theory proceeds from this insight. The area of mathematics and computer science called formal language theory provides a mathematical machinery for dealing with such phenomena.

What is language? An alphabet is a set containing a finite number of symbols. Possible alphabets for natural languages are the set of all phonemes or the set of all words of a language. For these two choices one obtains formal languages on different levels, but the mathematical principles are the same. Without loss of generality, we can consider the binary alphabet, {0,1}, by enumerating the actual alphabet in binary code.

A sentence is defined as a string of symbols. The set of all sentences over the binary alphabet is {0,1,00,01,10,11,000,...}. There are infinitely many sentences, as many as integers; the set of all sentences is 'countable'.

A language is a set of sentences. Among all possible sentences some are part of the language and some are not. A finite language contains a finite number of sentences. An infinite language contains an infinite number of sentences. There are infinitely many finite languages, as many as integers. There are infinitely many infinite languages, as many as real numbers; they are not countable. Hence, the set of all languages is not countable.

What is grammar? A grammar is a finite list of rules specifying a language. A grammar is expressed in terms of 'rewrite rules': a certain string can be rewritten as another string. Strings contain elements of the alphabet together with 'non-terminals', which are place holders. After iterated application of the rewrite rules, the final string will only contain symbols of the alphabet. Figures 1 and 2 give examples of grammars.


Figure 1 The basic objects of formal language theory are alphabets, sentences, languages and grammars. Full legend

High resolution image and legend (68k)




Figure 2 Three grammars and their corresponding languages. Full legend

High resolution image and legend (35k)



There are countably infinitely many grammars; any finite list of rewrite rules can be encoded by an integer. As there are uncountably many languages, only a small subset of them can be described by a grammar. These languages are called 'computable'.

Languages, grammars and machines There is a correspondence between languages, grammars and machines. 'Regular' languages are generated by finite-state grammars, which are equivalent to finite-state automata. Finite-state automata have a start, a finite number of intermediate states and a finish. Whenever the machine jumps from one state to the next, it emits an element of the alphabet. A particular run from start to finish produces a sentence. There are many different runs from start to finish, hence there are many different sentences. If a finite-state machine contains at least one loop, then it can generate infinitely many sentences. Finite-state automata accept all finite languages and some infinite languages.

'Context-free' languages are generated by context-free grammars, which can be implemented by push-down automata. These are computers with a single memory stack: at any one time they have only access to the top register of their memory.

'Context-sensitive' languages are generated by context-sensitive grammars. For each of these languages there exists a Turing machine, which can decide for every sentence whether it is part of the language or not. A Turing machine embodies the theoretical concept of a digital computer with infinite memory.

Computable languages are described by 'phrase structure' grammars that have unrestricted rewrite rules. For each computable language, there exists a Turing machine that can identify every sentence that is part of the language. If, however, the Turing machine receives as input a sentence which does not belong to the language, then it might compute forever. Hence, the Turing machine cannot always decide whether a sentence is part of the language or not.

Figure 3 shows the Chomsky hierarchy: finite-state grammars are a subset of context-free grammars, which are a subset of context-sensitive grammars, which are a subset of phrase-structure grammars, which are Turing complete.


Figure 3 The Chomsky hierarchy and the logical necessity of universal grammar. Full legend

High resolution image and legend (77k)



The structure of natural languages With the introduction of the Chomsky hierarchy, there was some interest in placing natural languages within this scheme. Natural languages are infinite: it is not possible to imagine a finite list that contains all English sentences. Furthermore, finite-state grammars are inadequate for natural languages. Such grammars are unable to represent long-range dependencies of the form 'if... then'. The string of words between 'if' and 'then' could be arbitrarily long, and could itself contain more paired if–then constructions. Such pairings relate to rules that generate strings of the form 0n1n, which require context-free grammars (Fig. 2). There is a continuing debate whether context-free grammars are adequate for natural languages, or whether more complex grammars need to be evoked29, 30.

The fundamental structures of natural languages are trees. The nodes represent phrases that can be composed of other phrases in a recursive manner. A tree is a 'derivation' of a sentence within the rule system of a particular grammar. The interpretation of a sentence depends on the underlying tree structure. Ambiguity arises if more than one tree can be associated with a given sentence. One can also define grammars that directly specify which trees are acceptable for a given language. Much of modern syntactic theory deals with such grammars31-34, which are, of course, also part of the Chomsky hierarchy, and the results of learning theory, to be discussed now, apply to them.

Learning theory
Learning is inductive inference. The learner is presented with data and has to infer the rules that generate these data. The difference between 'learning' and 'memorization' is the ability to generalize beyond one's own experience to novel circumstances. In the context of language, the child will generalize to novel sentences never heard before. Any person can produce and understand sentences that are not part of his previous linguistic experience. Learning theory describes the mathematics of learning with the aim of outlining conditions for successful generalization.

The paradox of language acquisition Children learn their native language by hearing grammatical sentences from their parents or others. From this 'environmental input', children construct an internal representation of the underlying grammar. Children are not told the grammatical rules. Neither children nor adults are ever aware of the grammatical rules that specify their own language.

Chomsky pointed out that the environmental input available to the child does not uniquely specify the grammatical rules35. This phenomenon is known as 'poverty of stimulus'36. 'The paradox of language acquisition' is that children of the same speech community reliably grow up to speak the same language37. The proposed solution is that children learn the correct grammar by choosing from a restricted set of candidate grammars. The 'theory' of this restricted set is 'universal grammar' (UG). Formally, UG is not a grammar, but a theory of a collection of grammars.

The concept of an innate, genetically determined UG was controversial when introduced some 40 years ago and has remained so. The mathematical approach of learning theory, however, can explain in what sense UG is a logical necessity.

Learnability Imagine a speaker–hearer pair. The speaker uses grammar, G, to construct sentences of language L. The hearer receives sentences and should after some time be able to use grammar G to construct other sentences of L. Mathematically speaking, the hearer is described by an algorithm (or more generally, a function), A, which takes a list of sentences as input and generates a language as output.

Let us introduce the notion of a 'text' as a list of sentences. Specifically, text T of language L is an infinite list of sentences of L with each sentence of L occurring at least once. Text TN contains the first N sentences of T. We say that language L is learnable by algorithm A if for each T of L there exists a number M such that for all N > M we have A(TN) = L. This means that, given enough sentences as input, the algorithm will provide the correct language as output.

Furthermore, a set of languages is learnable by an algorithm if each language of this set is learnable. We are interested in what set of languages, L = {L1,L2,..}, can be learned by a given algorithm.

A key result of learning theory, Gold's theorem23, implies there exists no algorithm that can learn the set of regular languages. As a consequence, no algorithm can learn a set of languages that contains the set of regular languages, such as the set of context-free languages, context-sensitive languages or computable languages.

Gold's theorem formally states there exists no algorithm that can learn a set of 'super-finite' languages. Such a set includes all finite languages and at least one infinite language. Intuitively, if the learner infers that the target language is an infinite language, whereas the actual target language is a finite language that is contained in the infinite language, then the learner will not encounter any contradicting evidence, and will never converge onto the correct language. This result holds in greatest possible generality: 'algorithm' here includes any function from text to language.

Probably almost correct A common criticism of Gold's framework is that the learner has to identify exactly the right language. For practical purposes, it might be sufficient that the learner acquires a grammar that is almost correct. Box 1 explains various extensions of the Gold framework, and in particular the approach of statistical learning theory. Here, the crucial requirement is that the learner converges with high probability to a language that is almost correct. Statistical learning theory also shows there is no procedure that can learn the set of all regular languages, thereby confirming the necessity of an innate UG. Some learning theories provide more information for the learner and thus allow larger classes of languages to be learnable, but no learning theory admits an unrestricted search space.

Learning finite languages Some readers might think that the arguments of learning theory rely on subtle properties of infinite languages. Let us therefore consider finite languages. In the context of statistical learning theory, the set of all finite languages cannot be learned. In the Gold framework, the set of all finite languages can be learned, but only by memorization: the learner will identify the correct language only after having heard all sentences of this language. A learner that considers the set of all finite languages has no possibility for generalization: the learner can never extrapolate beyond the sentences he has already encountered. This is not the case for natural language acquisition: we can always say new sentences.

Let us consider a finite set of finite languages. Suppose there are 3 sentences, S1, S2, S3. Hence there are 8 possible languages. Suppose learner A considers all 8 languages, while learner B considers only 2 languages, for example L1 = {S1, S2} and L2 = {S3}. If learner A receives sentence S1, he has no information whether sentences S2 or S3 will be part of the target language or not. He can only identify the target language after having heard all sentences. If learner B receives sentence S1 he knows that S2 will be part of the language, whereas S3 will not. He can extrapolate beyond his experience. The ability to search for underlying rules requires a restricted search space.

The necessity of innate expectations We can now state in what sense there has to be an innate UG. The human brain is equipped with a learning algorithm, AH, which enables us to learn certain languages. This algorithm can learn each of the existing 6,000 human languages and presumably many more, but it is impossible that AH could learn every computable language. Hence, there is a restricted set of languages that can be learned by AH. UG is the theory of this restricted set.

Learning theory suggests that a restricted search space has to exist before data. By 'data' we mean linguistic or other information the child uses to learn language or modify its language acquisition procedure. Therefore, in our terminology, 'before data' is equivalent to 'innate'. In this sense, learning theory shows there must be an innate UG, which is a consequence of the particular learning algorithm, AH, used by humans. Discovering properties of AH requires the empirical study of neurobiological and cognitive functions of the human brain involved in language acquisition. Some aspects of UG, however, might be unveiled by studying common features of existing human languages. This has been a major goal of linguistic research during the past decades. A particular approach is the 'principles and parameters theory', which assumes that the child comes equipped with innate principles and has to set parameters that are specific for individual languages38, 39. Another approach is 'optimality theory', where learning a specific language is ordering innate constraints40.

There is some discourse as to whether the learning mechanism, AH, is language specific or general purpose41. Ultimately this is a question about the particular architecture of the brain and which neurons participate in which computations, but one cannot deny that there is a learning mechanism, AH, that operates on linguistic input and enables the child to learn the rules of human language. This mechanism can learn a restricted set of languages; the theory of this set is UG. The continuing debate around an innate UG should not be whether there is one, but what form it takes41-43. One can dispute individual linguistic universals44, 45, but one cannot generally deny their existence.

Neural networks are an important tool for modelling the neural mechanisms of language acquisition. The results of learning theory also apply to neural networks: no neural network can learn an unrestricted set of languages46.

Sometimes it is claimed that the logical arguments for an innate UG rest on particular mathematical assumptions of generative grammars that deal only with syntax and not with semantics. Cognitive47, 48 and functional linguistics49 are not based on formal language theory, but use psychological objects such as symbols, categories, schemas and images. This does not remove the necessity of innate restrictions. The results of learning theory apply to any learning process, where a 'rule' has to be learned from some examples. Generalization is an inherent feature of any model of language acquisition, and applies to semantics, syntax and phonetics. Any procedure for successful generalization has to choose from a restricted range of hypotheses.

The results of learning theory also apply to learning mappings between linguistic form and meaning. If meaning is to be explicitly considered, then a language is not a set of sentences, but a set of sentence-meaning pairs (Box 2). The task of language acquisition is then to learn grammars that generate sentence-meaning pairs. Such grammars are also part of the Chomsky hierarchy, and there exists no learning procedure that can succeed on an unrestricted set of such languages.

What is special about language acquisition? Usually when we learn the grammar of generative systems, such as chess or arithmetic, somebody tells us the rules. We do not have to guess the moves of chess by looking at chess games. In contrast, the process of language acquisition occurs without being instructed about rules; neither teachers nor learners are aware of the rules. This is an important difference: if the learner is told the grammar of a language, then the set of all computable languages is learnable by an algorithm that memorizes the rules.

Evolutionary language theory
Humans and chimpanzees separated some 5 million years ago. Chimpanzees have a complex system of conceptual understanding and rich social interactions, but they do not have communication comparable to human language. The central question of the origin of human language is which genetic modifications led to changes in brain structures that were decisive for human language. Given the enormous complexity of this trait, we should expect several incremental steps guided by natural selection. In this process, evolution will have reused cognitive features that evolved long ago and for other purposes.

Understanding language evolution requires a theoretical framework explaining how darwinian dynamics lead to fundamental properties of human language such as arbitrary signs, lexicons, syntax and grammar50-62. Here we outline a minimalist program combining formal language theory, learning theory and evolutionary theory.

The basic approach is similar to evolutionary game theory. There is a population of individuals. Each individual uses a particular language. Individuals talk to each other. Successful communication results in a pay-off that contributes to fitness. Offspring inherit—subject to mutations—a mechanism to learn language and a UG. They use this mechanism to learn—subject to mistakes—the language of their parents or others.

Cultural evolution of language with constant universal grammar From a biological perspective, language is not the property of an individual, but the extended phenotype of a population. Let us consider a population where all individuals have the same UG, and let us assume that UG does not change from one generation to the next. Suppose that (in accordance with principles and parameters theory or optimality theory) UG specifies a finite number of languages L1,..Ln. Each individual in the population will speak one of those n languages.

We need a model of language that allows us to calculate the communicative pay-off, Fij, for an individual using Li talking to an individual using Lj. Box 2 outlines a fairly general approach, based on the assumption that a language can be seen as an infinite binary matrix linking phonetic forms to semantic forms. The languages, Li, can differ in their intrinsic communicative pay-off, Fii, which depends on ambiguity and expressive power. Some languages could be finite, others infinite.

Denote by xi the relative abundance of speakers of Li. The fitness of Li is given by fi = j = 1nxjFij. Individuals leave offspring proportional to their pay-off. The probability that a child will develop Lj if the parent uses Li is given by Qij. The 'learning matrix', Q, depends on the learning algorithm and UG. The language dynamical equation62 is given by:


The term -(x)xj ensures that ixi = 1. The variable (x) = ifi(x)xi denotes the average fitness of the population, and is a measure for linguistic coherence. The dynamics can also be interpreted in a purely cultural sense: individuals that communicate successfully are more likely to influence language acquisition of others. Equation (1) describes selection of languages for increased communicative function, Fii, and increased learnability, Qii.

For low accuracy of language acquisition, when Q is far from the identity matrix, there is no predominating language in the population, and the linguistic coherence is low. As the accuracy of language acquisition increases, and Q gets closer to the identity matrix, equilibrium solutions arise where a particular language is more abundant than others. The population has achieved linguistic coherence. The 'coherence threshold' specifies the minimum specificity of UG that is required for linguistic coherence (Fig. 4).


Figure 4 Linguistic coherence evolves if universal grammar (UG) is sufficiently specific. Full legend

High resolution image and legend (23k)



For certain learning mechanisms we can calculate the coherence threshold. The 'memoryless learner' starts with a randomly chosen language and stays with it as long as the input sentences are compatible with this language. If a sentence arrives that is not compatible, then the learner picks at random another language. The learner does not memorize which languages have already been rejected. The process stops after N sentences. Another mechanism is the 'batch learner', which memorizes N sentences, and at the end chooses the language that is most compatible with all N sentences.

If the similarity coefficients between languages, sij (Box 2), are constant, sij = s and sii = 1, then the memoryless learner has a coherence threshold N > C1n, whereas the batch learner has a coherence threshold N > C2 log n. If the sij values are taken from a uniform random distribution on the interval [0, 1] and if sii = 1, then the memoryless learner has a coherence threshold N > C3n log n, whereas the batch learner has a coherence threshold N > C4n (refs 63, 64). C1 to C4 are some constants. These conditions provide boundaries for the actual learning mechanism used by humans, which is arguably better than the memoryless learner and worse than the batch learner. The coherence threshold relates a life-history parameter of humans, N, to the maximum size of the search space, n, of UG.

Evolution of universal grammar Evolution of UG requires variation of UG. (UG is in fact neither a grammar nor universal.) Imagine a population of individuals using UGs U1 to UM. Each UI admits a subset of n grammars and determines a particular learning matrix QI. UI mutates genetically to UJ with probability WIJ. Deterministic population dynamics are given by:



This equation describes mutation and selection among M different universal grammars. The relative abundance of individuals with UG UJ speaking language Lj is given by xJj. At present little is known about the behaviour of this system. In the limit of no mutation among UGs, WII = 1, we find that the selective dynamics often lead to the elimination of all but one UG, but sometimes coexistence of different UGs can be observed. Equation (2) describes two processes on different timescales: the biological evolution of UG and the cultural evolution of language (Fig. 5).


Figure 5 Two aspects of language evolution. Full legend

High resolution image and legend (31k)



The ability to induce a coherent language is a major selective criterion for UG. A UG that induces linguistic coherence allows language adaptation and can be selected for linguistic function. There is also a trade-off between learnability and adaptability: a small search space (small n) is more likely to lead to linguistic coherence, but might exclude languages with high communicative pay-off.

Because the necessity of a restricted search space applies to any learning task, we can use an extended concept of UG for animal communication. During primate evolution, there was a succession of UGs that finally led to the UG of currently living humans. At some point a UG emerged that allowed languages of unlimited expressibility. Such evolutionary dynamics are described by equation (2).

Historical linguistics The language dynamical equation (1) can be used to study language change in the context of historical linguistics65-68. Languages change because the transmission from one generation to the next is not perfect. UG limits the type of variation that can occur. In the context of the principles-and-parameters theory, changes in syntax arise because children acquire different parameter settings66. Grammaticalization69 is the process where lexical items take on grammatical function. Creolization is the formation of a new language by children receiving mixed input70, 71. All such language changes can be studied mathematically. Many language changes are selectively neutral. Hence, we can use a neutral version of our approach possibly in conjunction with small population sizes and stochastic and spatial population dynamics. These are open problems.

Outlook
We have reviewed mathematical descriptions of language on three different levels: formal language theory, learning theory and evolution. These approaches need to be combined: ideas of language should be discussed in the context of acquisition, and ideas of acquisition in the context of evolution.

Some theoretical questions are: what is the interplay between the biological evolution of UG and the cultural evolution of language? What is the mechanism for adaptation among the various languages generated by a given UG? In terms of the principles-and-parameters theory, can we estimate the maximum number of parameters compatible with the coherence threshold? Some empirical questions are: what is the actual language learning algorithm used by humans? What are the restrictions imposed by UG? Can we identify genes that are crucial for linguistic or other cognitive functions? What can we say about the evolution of those genes?

The study of language as a biological phenomenon will bring together people from many disciplines including linguistics, cognitive science, psychology, genetics, animal behaviour, evolutionary biology, neurobiology and computer science. Fortunately we have language to talk to each other.



References 1. Pinker, S. & Bloom, A. Natural language and natural selection. Behav. Brain Sci. 13, 707-784 (1990) | ISI |
2. Jackendoff, R. Possible stages in the evolution of the language capacity. Trends Cogn. Sci. 3, 272-279 (1999) | PubMed | ISI |
3. Bickerton, D. Language and Species (Univ. Chicago Press, Chicago, 1990)
4. Lightfoot, D. The Development of Language: Acquisition, Changes and Evolution (Blackwell, Oxford, 1999).
5. Brandon, R. & Hornstein, N. From icon to symbol: Some speculations on the evolution of natural language. Phil. Biol. 1, 169-189 (1986)
6. Hurford, J. R., Studdert-Kennedy, M. A. & Knight, C. (eds) Approaches to the Evolution of Language (Cambridge Univ. Press, Cambridge, UK, 1998)
7. Newmeyer, F. Functional explanation in linguistics and the origins of language. Lang. Commun. 11, 3-28, 97-108 (1991) | ISI |
8. Lieberman, P. The Biology and Evolution of Language (Harvard Univ. Press, Cambridge, Massachusetts, 1984)
9. Maynard Smith, J. & Szathmary, E. The Major Transitions in Evolution (Freeman Spektrum, Oxford, 1995)
10. Hawkins, J. A. & Gell-Mann, M. The Evolution of Human Languages (Addison-Wesley, Reading, Massachusetts, 1992)
11. Aitchinson, J. The Seeds of Speech (Cambridge Univ. Press, Cambridge, UK, 1996)
12. Cavalli-Sforza, L. L. Genes, peoples and languages. Proc. Natl Acad. Sci. USA 94, 7719-7724 (1997) | PubMed | ISI |
13. Gopnik, M. & Crago, M. Familial aggregation of a developmental language disorder. Cognition 39, 1-50 (1991) | PubMed | ISI |
14. Lai, C. S. L., Fisher, S. E., Hurst, J. A., Vargha-Khadem, F. & Monaco, A. P. A forhead-domain gene is mutated in a severe speech and language disorder. Nature 413, 519-523 (2001) | Article | PubMed | ISI |
15. Deacon, T. The Symbolic Species (Penguin, London, 1997)
16. Vargha-Khadem, F. et al. Neural basis of an inherited speech and language disorder. Proc. Natl Acad. Sci. USA 95, 12695-12700 (1998) | Article | PubMed | ISI |
17. Smith, W. J. The Behaviour of Communicating (Harvard Univ. Press, Cambridge, UK, 1977)
18. Dunbar, R. Grooming, Gossip, and the Evolution of Language (Cambridge Univ. Press, Cambridge, UK, 1996)
19. Fitch, W. T. The evolution of speech: a comparative review. Trends Cogn. Sci. 4, 258-267 (2000) | PubMed | ISI |
20. Hauser, M. D. The Evolution of Communication (Harvard Univ. Press, Cambridge, Massachusetts, 1996)
21. Chomsky, N. A. Syntactic Structures (Mouton, New York, 1957)
22. Harrison, M. A. Introduction to Formal Language Theory (Addison-Wesley, Reading, Massachusetts, 1978)
23. Gold, E. M. Language identification in the limit. Informat. Control 10, 447-474 (1967) | ISI |
24. Vapnik, V. N. & Chervonenkis, A. Y. On the uniform convergence of relative frequencies of events to their probabilities. Theor. Prob. Applicat. 17, 264-280 (1971)
25. Valiant, L. G. A theory of learnable. Commun. ACM 27, 436-445 (1984)
26. Vapnik, V. N. Statistical Learning Theory (Wiley, New York, 1998)
27. Osherson, D., Stob, M. & Weinstein, S. Systems That Learn (MIT Press, Cambridge, Massachusetts, 1986)
28. Pinker, S. Formal models of language learning. Cognition 7, 217-283 (1979) | PubMed | ISI |
29. Pullum, G. K. & Gazdar, G. Natural languages and context free languages. Linguist. Phil. 4, 471-504 (1982) | ISI |
30. Shieber, S. M. Evidence against the context-freeness of natural language. Linguist. Phil. 8, 333-343 (1985) | ISI |
31. Chomsky, N. A. Lectures on Government and Binding: The Pisa Lectures (Foris, Dordrecht, 1984)
32. Sadock, J. M. Autolexical Syntax: A Theory of Parallel Grammatical Representations. Studies in Contemporary Linguistics (Univ. Chicago Press, Chicago, 1991)
33. Bresnan, J. Lexical-Functional Syntax (Blackwells, London, 2001)
34. Pollard, C. J. & Sag, I. A. Head-Driven Phrase Structure Grammar (Univ. Chicago Press, Chicago, 1994)
35. Chomsky, N. Language and Mind (Harcourt Brace Jovanovich, New York, 1972)
36. Wexler, K. & Culicover, P. Formal Principles of Language Acquisition (MIT Press, Cambridge, Massachusetts, 1980)
37. Jackendoff, R. Foundations of Language (Oxford Univ. Press, Oxford, 2001)
38. Chomsky, N. in Explanation in Linguistics (eds Hornstein, N. & Lightfoot, D.) 123-146 (Longman, London, 1981)
39. Baker, M. C. Atoms of Language (Basic Books, New York, 2001)
40. Prince, A. & Smolensky, P. Optimality: From neural networks to universal grammar. Science 275, 1604-1610 (1997) | PubMed | ISI |
41. Elman, J. L. Rethinking Innateness (MIT Press, Cambridge, Massachusetts, 1996)
42. Tomasello, M. The Cultural Origins of Human Cognition (Harvard Univ. Press, Cambridge, Massachusetts, 1999)
43. Sampson, G. Educating Eve: The Language Instinct Debate (Cassell Academic, London, 1999)
44. Greenberg, J. H., Ferguson, C. A. & Moravcsik, E. A. (eds) Universals of Human Language (Stanford Univ. Press, Stanford, 1978)
45. Comrie, B. Language Universals and Linguistic Typology (Univ. Chicago Press, Chicago, 1981)
46. Geman, S., Bienenstock, E. & Doursat, R. Neural networks and the bias/variance dilemma. Neural Comput. 4, 1-58 (1992) | ISI |
47. Langacker, R. Foundations of Cognitive Linguistics Vol. 1 (Stanford Univ. Press, Stanford, 1987)
48. Lakoff, G. Women, Fire and Dangerous Things: What Categories Reveal about the Mind (Univ. Chicago Press, Chicago, 1987)
49. Bates, E. & MacWhinney, B. Language Acquisition: The State of the Art (Cambridge Univ. Press, Cambridge, 1982)
50. Aoki, K. & Feldman, M. W. Toward a theory for the evolution of cultural communication: Coevolution of signal transmission and reception. Proc. Natl Acad. Sci. USA 84, 7164-7168 (1987) | PubMed | ISI |
51. Hurford, J. R. Biological evolution of the Saussurean sign as a component of the language acquisition device. Lingua 77, 187-222 (1989) | ISI |
52. Cangelosi, A. & Parisi, D. Simulating the Evolution of Language (Springer, London, 2002)
53. Kirby, S. & Hurford, J. Proc. Fourth European Conf. on Artificial Life (eds Husbands, P. & Harvey, I.) 493-502 (MIT Press, Cambridge, Massachusetts, 1997)
54. Steels, L. Proc. Fifth Artificial Life Conf. (eds Langton, C. G. & Shimohara, T.) 113-131 (MIT Press, Tokyo, 1996)
55. Nowak, M. A. & Krakauer, D. C. The evolution of language. Proc. Natl Acad. Sci. USA 96, 8028-8033 (1999) | Article | PubMed | ISI |
56. Nowak, M. A., Plotkin, J. B. & Jansen, V. A. A. Evolution of syntactic communication. Nature 404, 495-498 (2000) | Article | PubMed | ISI |
57. Komarova, N. L. & Nowak, M. A. Evolutionary dynamics of the lexical matrix. Bull. Math. Biol. 63, 451-485 (2001) | Article | PubMed | ISI |
58. Christiansen, M. H., Dale, R. A. C., Ellefson, M. R. & Conway, C. M. in Simulating the Evolution of Language (eds Cangelosi, A. & Parisi, D.) 165-187 (Springer, London, 2002)
59. Hashimoto, T. & Ikegami, T. Emergence of net-grammar in communicating agents. Biosystems 38, 1-14 (1996) | Article | PubMed | ISI |
60. Hazlehurst, B. & Hutchins, E. The emergence of propositions from the coordination of talk and action in a shared worlds. Lang. Cogn. Process. 13, 373-424 (1998) | ISI |
61. Pinker, S. The Language Instinct (Morrow, New York, 1994)
62. Nowak, M. A., Komarova, N. L. & Niyogi, P. Evolution of universal grammar. Science 291, 114-118 (2001) | PubMed | ISI |
63. Komarova, N. L. & Rivin, I. Mathematics of learning. Preprint math.PR/0105235 at http://lanl.arXiv.org (2001).
64. Rivin, I. Yet another zeta function and learning. Preprint cs.LG/0107033 at http://lanl.arXiv.org (2001).
65. Lightfoot, D. How to Set Parameters: Arguments from Language Change (MIT Press, Cambridge, Massachusetts, 1991)
66. Kroch, A. Reflexes of grammar in patterns of language change. Lang. Variat. Change 1, 199-244 (1989)
67. Wang, W. S. Y. in The Origins and Past of Modern Humans (eds Omoto, K. & Tobias, P. V.) 247-262 (World Scientific, Singapore, 1998)
68. Niyogi, P. & Berwick, R. C. Evolutionary consequences of language learning. Linguist. Phil. 20, 697-719 (1997) | ISI |
69. Hopper, P. & Traugott, E. Grammaticalization (Cambridge Univ. Press, Cambridge, 1993)
70. de Graff, M. Language Creation and Language Change: Creolization, Diachrony and Development (MIT Press, Cambridge, MA, 1999)
71. Mufwene, S. The Ecology of Language Evolution (Cambridge Univ. Press, Cambridge, 2001)
72. Angluin, D. Learning regular sets from queries and counterexamples. Informat. Comput. 75, 87-106 (1987) | ISI |
73. Angluin, D. & Kharitonov, M. When won't membership queries help? J. Comput. Syst. Sci. 50, 336-355 (1995) | ISI |
74. Gasarch, W. & Smith, C. Learning via queries. J. Assoc. Comput. Machin. 39, 649-674 (1992) | ISI |


Acknowledgements. Support from the David and Lucille Packard foundation, the Leon Levy and Shelby White initiatives fund, the Florence Gould foundation, the Ambrose Monell foundation, the National Science Foundation and J. E. Epstein is acknowledged.
Add a quick comment
Title
Your name Your email

Comment

Text Format
To add more detailed comments, or to upload files, see the full comment form.