Probabilistic context-free grammar

Last updated

Grammar theory to model symbol strings originated from work in computational linguistics aiming to understand the structure of natural languages. [1] [2] [3] Probabilistic context free grammars (PCFGs) have been applied in probabilistic modeling of RNA structures almost 40 years after they were introduced in computational linguistics. [4] [5] [6] [7] [8]

Contents

PCFGs extend context-free grammars similar to how hidden Markov models extend regular grammars. Each production is assigned a probability. The probability of a derivation (parse) is the product of the probabilities of the productions used in that derivation. These probabilities can be viewed as parameters of the model, and for large problems it is convenient to learn these parameters via machine learning. A probabilistic grammar's validity is constrained by context of its training dataset.

PCFGs have application in areas as diverse as natural language processing to the study the structure of RNA molecules and design of programming languages. Designing efficient PCFGs has to weigh factors of scalability and generality. Issues such as grammar ambiguity must be resolved. The grammar design affects results accuracy. Grammar parsing algorithms have various time and memory requirements.

Definitions

Derivation: The process of recursive generation of strings from a grammar.

Parsing: Finding a valid derivation using an automaton.

Parse Tree: The alignment of the grammar to a sequence.

An example of a parser for PCFG grammars is the pushdown automaton. The algorithm parses grammar nonterminals from left to right in a stack-like manner. This brute-force approach is not very efficient. In RNA secondary structure prediction variants of the Cocke–Younger–Kasami (CYK) algorithm provide more efficient alternatives to grammar parsing than pushdown automata. [9] Another example of a PCFG parser is the Stanford Statistical Parser which has been trained using Treebank. [10]

Formal definition

Similar to a CFG, a probabilistic context-free grammar G can be defined by a quintuple:

where

Relation with hidden Markov models

PCFGs models extend context-free grammars the same way as hidden Markov models extend regular grammars.

The Inside-Outside algorithm is an analogue of the Forward-Backward algorithm. It computes the total probability of all derivations that are consistent with a given sequence, based on some PCFG. This is equivalent to the probability of the PCFG generating the sequence, and is intuitively a measure of how consistent the sequence is with the given grammar. The Inside-Outside algorithm is used in model parametrization to estimate prior frequencies observed from training sequences in the case of RNAs.

Dynamic programming variants of the CYK algorithm find the Viterbi parse of a RNA sequence for a PCFG model. This parse is the most likely derivation of the sequence by the given PCFG.

Grammar construction

Context-free grammars are represented as a set of rules inspired from attempts to model natural languages. [3] The rules are absolute and have a typical syntax representation known as Backus–Naur form. The production rules consist of terminal and non-terminal S symbols and a blank may also be used as an end point. In the production rules of CFG and PCFG the left side has only one nonterminal whereas the right side can be any string of terminal or nonterminals. In PCFG nulls are excluded. [9] An example of a grammar:

This grammar can be shortened using the '|' ('or') character into:

Terminals in a grammar are words and through the grammar rules a non-terminal symbol is transformed into a string of either terminals and/or non-terminals. The above grammar is read as "beginning from a non-terminal S the emission can generate either a or b or ". Its derivation is:

Ambiguous grammar may result in ambiguous parsing if applied on homographs since the same word sequence can have more than one interpretation. Pun sentences such as the newspaper headline "Iraqi Head Seeks Arms" are an example of ambiguous parses.

One strategy of dealing with ambiguous parses (originating with grammarians as early as Pāṇini) is to add yet more rules, or prioritize them so that one rule takes precedence over others. This, however, has the drawback of proliferating the rules, often to the point where they become difficult to manage. Another difficulty is overgeneration, where unlicensed structures are also generated.

Probabilistic grammars circumvent these problems by ranking various productions on frequency weights, resulting in a "most likely" (winner-take-all) interpretation. As usage patterns are altered in diachronic shifts, these probabilistic rules can be re-learned, thus updating the grammar.

Assigning probability to production rules makes a PCFG. These probabilities are informed by observing distributions on a training set of similar composition to the language to be modeled. On most samples of broad language, probabilistic grammars where probabilities are estimated from data typically outperform hand-crafted grammars. CFGs when contrasted with PCFGs are not applicable to RNA structure prediction because while they incorporate sequence-structure relationship they lack the scoring metrics that reveal a sequence structural potential [11]

Weighted context-free grammar

A weighted context-free grammar (WCFG) is a more general category of context-free grammar, where each production has a numeric weight associated with it. The weight of a specific parse tree in a WCFG is the product [12] (or sum [13] ) of all rule weights in the tree. Each rule weight is included as often as the rule is used in the tree. A special case of WCFGs are PCFGs, where the weights are (logarithms of [14] [15] ) probabilities.

An extended version of the CYK algorithm can be used to find the "lightest" (least-weight) derivation of a string given some WCFG.

When the tree weight is the product of the rule weights, WCFGs and PCFGs can express the same set of probability distributions. [12]

Applications

RNA structure prediction

Energy minimization [16] [17] and PCFG provide ways of predicting RNA secondary structure with comparable performance. [4] [5] [9] However structure prediction by PCFGs is scored probabilistically rather than by minimum free energy calculation. PCFG model parameters are directly derived from frequencies of different features observed in databases of RNA structures [11] rather than by experimental determination as is the case with energy minimization methods. [18] [19]

The types of various structure that can be modeled by a PCFG include long range interactions, pairwise structure and other nested structures. However, pseudoknots can not be modeled. [4] [5] [9] PCFGs extend CFG by assigning probabilities to each production rule. A maximum probability parse tree from the grammar implies a maximum probability structure. Since RNAs preserve their structures over their primary sequence, RNA structure prediction can be guided by combining evolutionary information from comparative sequence analysis with biophysical knowledge about a structure plausibility based on such probabilities. Also search results for structural homologs using PCFG rules are scored according to PCFG derivations probabilities. Therefore, building grammar to model the behavior of base-pairs and single-stranded regions starts with exploring features of structural multiple sequence alignment of related RNAs. [9]

The above grammar generates a string in an outside-in fashion, that is the basepair on the furthest extremes of the terminal is derived first. So a string such as is derived by first generating the distal a's on both sides before moving inwards:

A PCFG model extendibility allows constraining structure prediction by incorporating expectations about different features of an RNA . Such expectation may reflect for example the propensity for assuming a certain structure by an RNA. [11] However incorporation of too much information may increase PCFG space and memory complexity and it is desirable that a PCFG-based model be as simple as possible. [11] [20]

Every possible string x a grammar generates is assigned a probability weight given the PCFG model . It follows that the sum of all probabilities to all possible grammar productions is . The scores for each paired and unpaired residue explain likelihood for secondary structure formations. Production rules also allow scoring loop lengths as well as the order of base pair stacking hence it is possible to explore the range of all possible generations including suboptimal structures from the grammar and accept or reject structures based on score thresholds. [9] [11]

Implementations

RNA secondary structure implementations based on PCFG approaches can be utilized in :

  • Finding consensus structure by optimizing structure joint probabilities over MSA. [20] [21]
  • Modeling base-pair covariation to detecting homology in database searches. [4]
  • pairwise simultaneous folding and alignment. [22] [23]

Different implementation of these approaches exist. For example, Pfold is used in secondary structure prediction from a group of related RNA sequences, [20] covariance models are used in searching databases for homologous sequences and RNA annotation and classification, [4] [24] RNApromo, CMFinder and TEISER are used in finding stable structural motifs in RNAs. [25] [26] [27]

Design considerations

PCFG design impacts the secondary structure prediction accuracy. Any useful structure prediction probabilistic model based on PCFG has to maintain simplicity without much compromise to prediction accuracy. Too complex a model of excellent performance on a single sequence may not scale. [9] A grammar based model should be able to:

  • Find the optimal alignment between a sequence and the PCFG.
  • Score the probability of the structures for the sequence and subsequences.
  • Parameterize the model by training on sequences/structures.
  • Find the optimal grammar parse tree (CYK algorithm).
  • Check for ambiguous grammar (Conditional Inside algorithm).

The resulting of multiple parse trees per grammar denotes grammar ambiguity. This may be useful in revealing all possible base-pair structures for a grammar. However an optimal structure is the one where there is one and only one correspondence between the parse tree and the secondary structure.

Two types of ambiguities can be distinguished. Parse tree ambiguity and structural ambiguity. Structural ambiguity does not affect thermodynamic approaches as the optimal structure selection is always on the basis of lowest free energy scores. [11] Parse tree ambiguity concerns the existence of multiple parse trees per sequence. Such an ambiguity can reveal all possible base-paired structures for the sequence by generating all possible parse trees then finding the optimal one. [28] [29] [30] In the case of structural ambiguity multiple parse trees describe the same secondary structure. This obscures the CYK algorithm decision on finding an optimal structure as the correspondence between the parse tree and the structure is not unique. [31] Grammar ambiguity can be checked for by the conditional-inside algorithm. [9] [11]

Building a PCFG model

A probabilistic context free grammar consists of terminal and nonterminal variables. Each feature to be modeled has a production rule that is assigned a probability estimated from a training set of RNA structures. Production rules are recursively applied until only terminal residues are left.

A starting non-terminal produces loops. The rest of the grammar proceeds with parameter that decide whether a loop is a start of a stem or a single stranded region s and parameter that produces paired bases.

The formalism of this simple PCFG looks like:

The application of PCFGs in predicting structures is a multi-step process. In addition, the PCFG itself can be incorporated into probabilistic models that consider RNA evolutionary history or search homologous sequences in databases. In an evolutionary history context inclusion of prior distributions of RNA structures of a structural alignment in the production rules of the PCFG facilitates good prediction accuracy. [21]

A summary of general steps for utilizing PCFGs in various scenarios:

  • Generate production rules for the sequences.
  • Check ambiguity.
  • Recursively generate parse trees of the possible structures using the grammar.
  • Rank and score the parse trees for the most plausible sequence. [9]

Algorithms

Several algorithms dealing with aspects of PCFG based probabilistic models in RNA structure prediction exist. For instance the inside-outside algorithm and the CYK algorithm. The inside-outside algorithm is a recursive dynamic programming scoring algorithm that can follow expectation-maximization paradigms. It computes the total probability of all derivations that are consistent with a given sequence, based on some PCFG. The inside part scores the subtrees from a parse tree and therefore subsequences probabilities given an PCFG. The outside part scores the probability of the complete parse tree for a full sequence. [32] [33] CYK modifies the inside-outside scoring. Note that the term 'CYK algorithm' describes the CYK variant of the inside algorithm that finds an optimal parse tree for a sequence using a PCFG. It extends the actual CYK algorithm used in non-probabilistic CFGs. [9]

The inside algorithm calculates probabilities for all of a parse subtree rooted at for subsequence . Outside algorithm calculates probabilities of a complete parse tree for sequence x from root excluding the calculation of . The variables α and β refine the estimation of probability parameters of an PCFG. It is possible to reestimate the PCFG algorithm by finding the expected number of times a state is used in a derivation through summing all the products of α and β divided by the probability for a sequence x given the model . It is also possible to find the expected number of times a production rule is used by an expectation-maximization that utilizes the values of α and β. [32] [33] The CYK algorithm calculates to find the most probable parse tree and yields . [9]

Memory and time complexity for general PCFG algorithms in RNA structure predictions are and respectively. Restricting a PCFG may alter this requirement as is the case with database searches methods.

Covariance models (CMs) are a special type of PCFGs with applications in database searches for homologs, annotation and RNA classification. Through CMs it is possible to build PCFG-based RNA profiles where related RNAs can be represented by a consensus secondary structure. [4] [5] The RNA analysis package Infernal uses such profiles in inference of RNA alignments. [34] The Rfam database also uses CMs in classifying RNAs into families based on their structure and sequence information. [24]

CMs are designed from a consensus RNA structure. A CM allows indels of unlimited length in the alignment. Terminals constitute states in the CM and the transition probabilities between the states is 1 if no indels are considered. [9] Grammars in a CM are as follows:

probabilities of pairwise interactions between 16 possible pairs
probabilities of generating 4 possible single bases on the left
probabilities of generating 4 possible single bases on the right
bifurcation with a probability of 1
start with a probability of 1
end with a probability of 1

The model has 6 possible states and each state grammar includes different types of secondary structure probabilities of the non-terminals. The states are connected by transitions. Ideally current node states connect to all insert states and subsequent node states connect to non-insert states. In order to allow insertion of more than one base insert states connect to themselves. [9]

In order to score a CM model the inside-outside algorithms are used. CMs use a slightly different implementation of CYK. Log-odds emission scores for the optimum parse tree - - are calculated out of the emitting states . Since these scores are a function of sequence length a more discriminative measure to recover an optimum parse tree probability score- - is reached by limiting the maximum length of the sequence to be aligned and calculating the log-odds relative to a null. The computation time of this step is linear to the database size and the algorithm has a memory complexity of . [9]

Example: Using evolutionary information to guide structure prediction

The KH-99 algorithm by Knudsen and Hein lays the basis of the Pfold approach to predicting RNA secondary structure. [20] In this approach the parameterization requires evolutionary history information derived from an alignment tree in addition to probabilities of columns and mutations. The grammar probabilities are observed from a training dataset.

Estimate column probabilities for paired and unpaired bases

In a structural alignment the probabilities of the unpaired bases columns and the paired bases columns are independent of other columns. By counting bases in single base positions and paired positions one obtains the frequencies of bases in loops and stems. For basepair X and Y an occurrence of is also counted as an occurrence of . Identical basepairs such as are counted twice.

Calculate mutation rates for paired and unpaired bases

By pairing sequences in all possible ways overall mutation rates are estimated. In order to recover plausible mutations a sequence identity threshold should be used so that the comparison is between similar sequences. This approach uses 85% identity threshold between pairing sequences. First single base positions differences -except for gapped columns- between sequence pairs are counted such that if the same position in two sequences had different bases X, Y the count of the difference is incremented for each sequence.

while  first sequence  pair second sequence pair
Calculate mutation rates.Let   mutation of base X to base Y Let   the negative of the rate of X mutation to other bases  the probability that the base is not paired.

For unpaired bases a 4 X 4 mutation rate matrix is used that satisfies that the mutation flow from X to Y is reversible: [35]

For basepairs a 16 X 16 rate distribution matrix is similarly generated. [36] [37] The PCFG is used to predict the prior probability distribution of the structure whereas posterior probabilities are estimated by the inside-outside algorithm and the most likely structure is found by the CYK algorithm. [20]

Estimate alignment probabilities

After calculating the column prior probabilities the alignment probability is estimated by summing over all possible secondary structures. Any column C in a secondary structure for a sequence D of length l such that can be scored with respect to the alignment tree T and the mutational model M. The prior distribution given by the PCFG is . The phylogenetic tree, T can be calculated from the model by maximum likelihood estimation. Note that gaps are treated as unknown bases and the summation can be done through dynamic programming. [38]

Assign production probabilities to each rule in the grammar

Each structure in the grammar is assigned production probabilities devised from the structures of the training dataset. These prior probabilities give weight to predictions accuracy. [21] [32] [33] The number of times each rule is used depends on the observations from the training dataset for that particular grammar feature. These probabilities are written in parentheses in the grammar formalism and each rule will have a total of 100%. [20] For instance:

Predict the structure likelihood

Given the prior alignment frequencies of the data the most likely structure from the ensemble predicted by the grammar can then be computed by maximizing through the CYK algorithm. The structure with the highest predicted number of correct predictions is reported as the consensus structure. [20]

Pfold improvements on the KH-99 algorithm

PCFG based approaches are desired to be scalable and general enough. Compromising speed for accuracy needs to as minimal as possible. Pfold addresses the limitations of the KH-99 algorithm with respect to scalability, gaps, speed and accuracy. [20]

  • In Pfold gaps are treated as unknown. In this sense the probability of a gapped column equals that of an ungapped one.
  • In Pfold the tree T is calculated prior to structure prediction through neighbor joining and not by maximum likelihood through the PCFG grammar. Only the branch lengths are adjusted to maximum likelihood estimates.
  • An assumption of Pfold is that all sequences have the same structure. Sequence identity threshold and allowing a 1% probability that any nucleotide becomes another limit the performance deterioration due to alignment errors.

Protein sequence analysis

Whereas PCFGs have proved powerful tools for predicting RNA secondary structure, usage in the field of protein sequence analysis has been limited. Indeed, the size of the amino acid alphabet and the variety of interactions seen in proteins make grammar inference much more challenging. [39] As a consequence, most applications of formal language theory to protein analysis have been mainly restricted to the production of grammars of lower expressive power to model simple functional patterns based on local interactions. [40] [41] Since protein structures commonly display higher-order dependencies including nested and crossing relationships, they clearly exceed the capabilities of any CFG. [39] Still, development of PCFGs allows expressing some of those dependencies and providing the ability to model a wider range of protein patterns.

See also

Related Research Articles

<span class="mw-page-title-main">Context-free grammar</span> Type of formal grammar

In formal language theory, a context-free grammar (CFG) is a formal grammar whose production rules can be applied to a nonterminal symbol regardless of its context. In particular, in a context-free grammar, each production rule is of the form

In formal language theory, a context-free language (CFL) is a language generated by a context-free grammar (CFG).

In computer science, the Earley parser is an algorithm for parsing strings that belong to a given context-free language, though it may suffer problems with certain nullable grammars. The algorithm, named after its inventor, Jay Earley, is a chart parser that uses dynamic programming; it is mainly used for parsing in computational linguistics. It was first introduced in his dissertation in 1968.

In computer science, the Cocke–Younger–Kasami algorithm is a parsing algorithm for context-free grammars published by Itiroo Sakai in 1961. The algorithm is named after some of its rediscoverers: John Cocke, Daniel Younger, Tadao Kasami, and Jacob T. Schwartz. It employs bottom-up parsing and dynamic programming.

Pattern recognition is the task of assigning a class to an observation based on patterns extracted from data. While similar, pattern recognition (PR) is not to be confused with pattern machines (PM) which may possess (PR) capabilities but their primary function is to distinguish and create emergent pattern. PR has applications in statistical data analysis, signal processing, image analysis, information retrieval, bioinformatics, data compression, computer graphics and machine learning. Pattern recognition has its origins in statistics and engineering; some modern approaches to pattern recognition include the use of machine learning, due to the increased availability of big data and a new abundance of processing power.

A Bayesian network is a probabilistic graphical model that represents a set of variables and their conditional dependencies via a directed acyclic graph (DAG). While it is one of several forms of causal notation, causal networks are special cases of Bayesian networks. Bayesian networks are ideal for taking an event that occurred and predicting the likelihood that any one of several possible known causes was the contributing factor. For example, a Bayesian network could represent the probabilistic relationships between diseases and symptoms. Given symptoms, the network can be used to compute the probabilities of the presence of various diseases.

Parsing, syntax analysis, or syntactic analysis is the process of analyzing a string of symbols, either in natural language, computer languages or data structures, conforming to the rules of a formal grammar. The term parsing comes from Latin pars (orationis), meaning part.

Top-down parsing in computer science is a parsing strategy where one first looks at the highest level of the parse tree and works down the parse tree by using the rewriting rules of a formal grammar. LL parsers are a type of parser that uses a top-down parsing strategy.

In computer science, an ambiguous grammar is a context-free grammar for which there exists a string that can have more than one leftmost derivation or parse tree. Every non-empty context-free language admits an ambiguous grammar by introducing e.g. a duplicate rule. A language that only admits ambiguous grammars is called an inherently ambiguous language. Deterministic context-free grammars are always unambiguous, and are an important subclass of unambiguous grammars; there are non-deterministic unambiguous grammars, however.

In computer science, a parsing expression grammar (PEG) is a type of analytic formal grammar, i.e. it describes a formal language in terms of a set of rules for recognizing strings in the language. The formalism was introduced by Bryan Ford in 2004 and is closely related to the family of top-down parsing languages introduced in the early 1970s. Syntactically, PEGs also look similar to context-free grammars (CFGs), but they have a different interpretation: the choice operator selects the first match in PEG, while it is ambiguous in CFG. This is closer to how string recognition tends to be done in practice, e.g. by a recursive descent parser.

In the formal language theory of computer science, left recursion is a special case of recursion where a string is recognized as part of a language by the fact that it decomposes into a string from that same language and a suffix. For instance, can be recognized as a sum because it can be broken into , also a sum, and , a suitable suffix.

Conditional random fields (CRFs) are a class of statistical modeling methods often applied in pattern recognition and machine learning and used for structured prediction. Whereas a classifier predicts a label for a single sample without considering "neighbouring" samples, a CRF can take context into account. To do so, the predictions are modelled as a graphical model, which represents the presence of dependencies between the predictions. What kind of graph is used depends on the application. For example, in natural language processing, "linear chain" CRFs are popular, for which each prediction is dependent only on its immediate neighbours. In image processing, the graph typically connects locations to nearby and/or similar locations to enforce that they receive similar predictions.

Nucleic acid structure prediction is a computational method to determine secondary and tertiary nucleic acid structure from its sequence. Secondary structure can be predicted from one or several nucleic acid sequences. Tertiary structure can be predicted from the sequence, or by comparative modeling.

In the mathematical theory of stochastic processes, variable-order Markov (VOM) models are an important class of models that extend the well known Markov chain models. In contrast to the Markov chain models, where each random variable in a sequence with a Markov property depends on a fixed number of random variables, in VOM models this number of conditioning random variables may vary based on the specific observed realization.

In bioinformatics, Stemloc is an open source software for multiple RNA sequence alignment and RNA structure prediction based on probabilistic models of RNA structure known as Pair stochastic context-free grammars. Stemloc attempts to simultaneously predict and align the structure of RNA sequences with an improved time and space cost compared to previous methods with the same motive. The resulting software implements constrained versions of the Sankoff algorithm by introducing both fold and alignment constraints, which reduces processor and memory usage and allows for larger RNA sequences to be analyzed on commodity hardware. Stemloc was written in 2004 by Ian Holmes.

In computer programming, a parser combinator is a higher-order function that accepts several parsers as input and returns a new parser as its output. In this context, a parser is a function accepting strings as input and returning some structure as output, typically a parse tree or a set of indices representing locations in the string where parsing stopped successfully. Parser combinators enable a recursive descent parsing strategy that facilitates modular piecewise construction and testing. This parsing technique is called combinatory parsing.

<span class="mw-page-title-main">Formal grammar</span> Structure of a formal language

A formal grammar describes how to form strings from an alphabet of a formal language that are valid according to the language's syntax. A grammar does not describe the meaning of the strings or what can be done with them in whatever context—only their form. A formal grammar is defined as a set of production rules for such strings in a formal language.

Structured prediction or structured (output) learning is an umbrella term for supervised machine learning techniques that involves predicting structured objects, rather than scalar discrete or real values.

Stochastic chains with memory of variable length are a family of stochastic chains of finite order in a finite alphabet, such as, for every time pass, only one finite suffix of the past, called context, is necessary to predict the next symbol. These models were introduced in the information theory literature by Jorma Rissanen in 1983, as a universal tool to data compression, but recently have been used to model data in different areas such as biology, linguistics and music.

Syntactic parsing is the automatic analysis of syntactic structure of natural language, especially syntactic relations and labelling spans of constituents. It is motivated by the problem of structural ambiguity in natural language: a sentence can be assigned multiple grammatical parses, so some kind of knowledge beyond computational grammar rules is needed to tell which parse is intended. Syntactic parsing is one of the important tasks in computational linguistics and natural language processing, and has been a subject of research since the mid-20th century with the advent of computers.

References

  1. Chomsky, Noam (1956). "Three models for the description of language". IRE Transactions on Information Theory. 2 (3): 113–124. doi:10.1109/TIT.1956.1056813. S2CID   19519474.
  2. Chomsky, Noam (June 1959). "On certain formal properties of grammars". Information and Control. 2 (2): 137–167. doi: 10.1016/S0019-9958(59)90362-6 .
  3. 1 2 Noam Chomsky, ed. (1957). Syntactic Structures. Mouton & Co. Publishers, Den Haag, Netherlands.
  4. 1 2 3 4 5 6 Eddy S. R. & Durbin R. (1994). "RNA sequence analysis using covariance models". Nucleic Acids Research. 22 (11): 2079–2088. doi:10.1093/nar/22.11.2079. PMC   308124 . PMID   8029015.
  5. 1 2 3 4 Sakakibara Y.; Brown M.; Hughey R.; Mian I. S.; et al. (1994). "Stochastic context-free grammars for tRNA modelling". Nucleic Acids Research. 22 (23): 5112–5120. doi:10.1093/nar/22.23.5112. PMC   523785 . PMID   7800507.
  6. Grat, L. (1995). "Automatic RNA secondary structure determination with stochastic context-free grammars" (PDF). In Rawlings, C., Clark, D., Altman, R., Hunter, L., Lengauer, T and Wodak, S. Proceedings of the Third International Conference on Intelligent Systems for Molecular Biology, AAAI Press: 136–144. Archived from the original (PDF) on 2015-12-04. Retrieved 2017-08-03.
  7. Lefebvre, F (1995). "An optimized parsing algorithm well suited to RNA folding". In Rawlings, C.; Clark, D.; Altman, R.; Hunter, L.; Lengauer, T.; Wodak, S. (eds.). Proceedings of the Third International Conference on Intelligent Systems for Molecular Biology (PDF). AAAI Press. pp. 222–230.
  8. Lefebvre, F. (1996). "A grammar-based unification of several alignment and folding algorithms". In States, D. J.; Agarwal, P.; Gaasterlan, T.; Hunter, L.; Smith R. F. (eds.). Proceedings of the Fourth International Conference on Intelligent Systems for Molecular Biology (PDF). AAAI Press. pp. 143–153.
  9. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 R. Durbin; S. Eddy; A. Krogh; G. Mitchinson (1998). Biological sequence analysis: probabilistic models of proteins and nucleic acids. Cambridge University Press. ISBN   978-0-521-62971-3.
  10. Klein, Daniel; Manning, Christopher (2003). "Accurate Unlexicalized Parsing" (PDF). Proceedings of the 41st Meeting of the Association for Computational Linguistics: 423–430.
  11. 1 2 3 4 5 6 7 Dowell R. & Eddy S. (2004). "Evaluation of several lightweight stochastic context-free grammars for RNA secondary structure prediction". BMC Bioinformatics. 5 (71): 71. doi: 10.1186/1471-2105-5-71 . PMC   442121 . PMID   15180907.
  12. 1 2 Smith, Noah A.; Johnson, Mark (2007). "Weighted and Probabilistic Context-Free Grammars Are Equally Expressive" (PDF). Computational Linguistics. 33 (4): 477. doi:10.1162/coli.2007.33.4.477. S2CID   1405777.
  13. Katsirelos, George; Narodytska, Nina; Walsh, Toby (2008). "The Weighted CFG Constraint". Integration of AI and OR Techniques in Constraint Programming for Combinatorial Optimization Problems. Lecture Notes in Computer Science. Vol. 5015. pp. 323–327. CiteSeerX   10.1.1.150.1187 . doi:10.1007/978-3-540-68155-7_31. ISBN   978-3-540-68154-0. S2CID   9375313.
  14. Johnson, Mark (2005). "log linear or Gibbs models" (PDF).
  15. Chi, Zhiyi (March 1999). "Statistical properties of probabilistic context-free grammars" (PDF). Computational Linguistics. 25 (1): 131–160. Archived from the original (PDF) on 2010-08-21.
  16. McCaskill J. S. (1990). "The Equilibrium Partition Function and Base Pair Binding Probabilities for RNA Secondary Structure". Biopolymers. 29 (6–7): 1105–19. doi:10.1002/bip.360290621. hdl: 11858/00-001M-0000-0013-0DE3-9 . PMID   1695107. S2CID   12629688.
  17. Juan V.; Wilson C. (1999). "RNA Secondary Structure Prediction Based on Free Energy and Phylogenetic Analysis". J. Mol. Biol. 289 (4): 935–947. doi:10.1006/jmbi.1999.2801. PMID   10369773.
  18. Zuker M (2000). "Calculating Nucleic Acid Secondary Structure". Curr. Opin. Struct. Biol. 10 (3): 303–310. doi:10.1016/S0959-440X(00)00088-9. PMID   10851192.
  19. Mathews D. H.; Sabina J.; Zuker M.; Turner D. H. (1999). "Expanded sequence dependence of thermodynamic parameters improves prediction of RNA secondary structure". J. Mol. Biol. 288 (5): 911–940. doi: 10.1006/jmbi.1999.2700 . PMID   10329189. S2CID   19989405.
  20. 1 2 3 4 5 6 7 8 B. Knudsen & J. Hein. (2003). "Pfold: RNA secondary structure prediction using stochastic context-free grammars". Nucleic Acids Research. 31 (13): 3423–3428. doi:10.1093/nar/gkg614. PMC   169020 . PMID   12824339.
  21. 1 2 3 Knudsen B.; Hein J. (1999). "RNA Secondary Structure Prediction Using Stochastic Context-Free Grammars and Evolutionary History". Bioinformatics. 15 (6): 446–454. doi: 10.1093/bioinformatics/15.6.446 . PMID   10383470.
  22. Rivas E.; Eddy S. R. (2001). "Noncoding RNA Gene Detection Using Comparative Sequence Analysis". BMC Bioinformatics. 2 (1): 8. doi: 10.1186/1471-2105-2-8 . PMC   64605 . PMID   11801179.
  23. Holmes I.; Rubin G. M. (2002). Pairwise RNA Structure Comparison with Stochastic Context-Free Grammars . pp.  163–174. doi:10.1142/9789812799623_0016. ISBN   978-981-02-4777-5. PMID   11928472.{{cite book}}: |journal= ignored (help)
  24. 1 2 P. P. Gardner; J. Daub; J. Tate; B. L. Moore; I. H. Osuch; S. Griffiths-Jones; R. D. Finn; E. P. Nawrocki; D. L. Kolbe; S. R. Eddy; A. Bateman. (2011). "Rfam: Wikipedia, clans and the "decimal" release". Nucleic Acids Research. 39 (Suppl 1): D141–D145. doi:10.1093/nar/gkq1129. PMC   3013711 . PMID   21062808.
  25. Yao Z.; Weinberg Z.; Ruzzo W. L. (2006). "CMfinder-a covariance model based RNA motif finding algorithm". Bioinformatics. 22 (4): 445–452. doi: 10.1093/bioinformatics/btk008 . PMID   16357030.
  26. Rabani M.; Kertesz M.; Segal E. (2008). "Computational prediction of RNA structural motifs involved in post-transcriptional regulatory processes". Proc. Natl. Acad. Sci. USA. 105 (39): 14885–14890. Bibcode:2008PNAS..10514885R. doi: 10.1073/pnas.0803169105 . PMC   2567462 . PMID   18815376.
  27. Goodarzi H.; Najafabadi H. S.; Oikonomou P.; Greco T. M.; Fish L.; Salavati R.; Cristea I. M.; Tavazoie S. (2012). "Systematic discovery of structural elements governing stability of mammalian messenger RNAs". Nature. 485 (7397): 264–268. Bibcode:2012Natur.485..264G. doi:10.1038/nature11013. PMC   3350620 . PMID   22495308.
  28. Sipser M. (1996). Introduction to Theory of Computation. Brooks Cole Pub Co.
  29. Michael A. Harrison (1978). Introduction to Formal Language Theory. Addison-Wesley.
  30. Hopcroft J. E.; Ullman J. D. (1979). Introduction to Automata Theory, Languages, and Computation. Addison-Wesley.
  31. Giegerich R. (2000). "Explaining and Controlling Ambiguity in Dynamic Programming". Combinatorial Pattern Matching. Lecture Notes in Computer Science. Vol. 1848. In Proceedings of the 11th Annual Symposium on Combinatorial Pattern Matching 1848 Edited by: Giancarlo R., Sankoff D. Montréal, Canada: Springer-Verlag, Berlin. pp. 46–59. doi:10.1007/3-540-45123-4_6. ISBN   978-3-540-67633-1. S2CID   17088251.
  32. 1 2 3 Lari K.; Young S. J. (1990). "The estimation of stochastic context-free grammars using the inside-outside algorithm". Computer Speech and Language. 4: 35–56. doi:10.1016/0885-2308(90)90022-X.
  33. 1 2 3 Lari K.; Young S. J. (1991). "Applications of stochastic context-free grammars using the inside-outside algorithm". Computer Speech and Language. 5 (3): 237–257. doi:10.1016/0885-2308(91)90009-F.
  34. Nawrocki E. P., Eddy S. R. (2013). "Infernal 1.1:100-fold faster RNA homology searches". Bioinformatics. 29 (22): 2933–2935. doi:10.1093/bioinformatics/btt509. PMC   3810854 . PMID   24008419.
  35. Tavaré S. (1986). "Some probabilistic and statistical problems in the analysis of DNA sequences". Lectures on Mathematics in the Life Sciences. American Mathematical Society. 17: 57–86.
  36. Muse S. V. (1995). "Evolutionary analyses of DNA sequences subject to constraints of secondary structure". Genetics. 139 (3): 1429–1439. doi:10.1093/genetics/139.3.1429. PMC   1206468 . PMID   7768450.
  37. Schöniger M.; von Haeseler A. (1994). "A stochastic model for the evolution of autocorrelated DNA sequences". Mol. Phylogenet. Evol. 3 (3): 240–7. doi:10.1006/mpev.1994.1026. PMID   7529616.
  38. Baker, J. K. (1979). "Trainable grammars for speech recognition". The Journal of the Acoustical Society of America. 65 (S1): S132. Bibcode:1979ASAJ...65Q.132B. doi: 10.1121/1.2017061 .
  39. 1 2 Searls, D (2013). "Review: A primer in macromolecular linguistics". Biopolymers. 99 (3): 203–217. doi:10.1002/bip.22101. PMID   23034580. S2CID   12676925.
  40. Krogh, A; Brown, M; Mian, I; Sjolander, K; Haussler, D (1994). "Hidden Markov models in computational biology: Applications to protein modeling". J Mol Biol. 235 (5): 1501–1531. doi:10.1006/jmbi.1994.1104. PMID   8107089. S2CID   2160404.
  41. Sigrist, C; Cerutti, L; Hulo, N; Gattiker, A; Falquet, L; Pagni, M; Bairoch, A; Bucher, P (2002). "PROSITE: a documented database using patterns and profiles as motif descriptors". Brief Bioinform. 3 (3): 265–274. doi: 10.1093/bib/3.3.265 . PMID   12230035.