Skip to main content
Log in

Dynamic Hybrid Random Fields for the Probabilistic Graphical Modeling of Sequential Data: Definitions, Algorithms, and an Application to Bioinformatics

  • Published:
Neural Processing Letters Aims and scope Submit manuscript

Abstract

The paper introduces a dynamic extension of the hybrid random field (HRF), called dynamic HRF (D-HRF). The D-HRF is aimed at the probabilistic graphical modeling of arbitrary-length sequences of sets of (time-dependent) discrete random variables under Markov assumptions. Suitable maximum likelihood algorithms for learning the parameters and the structure of the D-HRF are presented. The D-HRF inherits the computational efficiency and the modeling capabilities of HRFs, subsuming both dynamic Bayesian networks and Markov random fields. The behavior of the D-HRF is first evaluated empirically on synthetic data drawn from probabilistic distributions having known form. Then, D-HRFs (combined with a recurrent autoencoder) are successfully applied to the prediction of the disulfide-bonding state of cysteines from the primary structure of proteins in the Protein Data Bank.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. Bearing in mind this sequential dynamics of the model, we will occasionally refer to the RNN as operating “over time”, such that it is fed with t-th residue at “time t”.

  2. Available publicly at http://www.biocomp.unibo.it/savojard/PDBCYS.ssbonds.txt.

References

  1. Baldi P, Brunak S (1998) Bioinformatics: the machine learning approach. MIT Press, Cambridge

    MATH  Google Scholar 

  2. Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE (2000) The protein data bank. Nucleic Acids Res 28(1):235–242

    Article  Google Scholar 

  3. Besag J (1975) Statistical analysis of non-lattice data. The Statistician 24:179–195

    Article  Google Scholar 

  4. Bongini M, Laveglia V, Trentin E (2016) A hybrid recurrent neural network/dynamic probabilistic graphical model predictor of the disulfide bonding state of cysteines from the primary structure of proteins. In: Proceedings of the 7th IAPR TC 3 workshop on artificial neural networks in pattern recognition, pp 257–268. Springer

  5. Bongini M, Trentin E (2012) Towards a novel probabilistic graphical model of sequential data: a solution to the problem of structure learning and an empirical evaluation. In: Proceedings of the 5th INNS IAPR TC 3 GIRPR workshop on artificial neural networks in pattern recognition, pp. 82–92. Springer

  6. Cappé O, Moulines E, Ryden T (2005) Inference in hidden Markov models. Springer, New York

    MATH  Google Scholar 

  7. Ceroni A, Passerini A, Vullo A, Frasconi P (2006) DISULFIND: a disulfide bonding state and cysteine connectivity prediction server. Nucleic Acids Res 34(Web–Server–Issue):177–181

    Article  Google Scholar 

  8. Chen Y, Lin Y, Lin C, Hwang J (2004) Prediction of the bonding states of cysteines using the support vector machines based on multiple feature vectors and cysteine state sequences. Proteins 55(4):1036–1042

    Article  Google Scholar 

  9. Chung W, Yang C, Hor C (2009) An effective tuning method for cysteine state classification. In: Proceedings of the national computer symposium, workshop on algorithms and bioinformatics, Taipei, Taiwan, 27–28 November

  10. Cooper GF, Herskovits E (1992) A Bayesian method for the induction of probabilistic networks from data. Mach Learn 9:309–347

    MATH  Google Scholar 

  11. Crick F (1970) Central dogma of molecular biology. Nature 227(5258):561–563

    Article  Google Scholar 

  12. Elidan G, Gould S (2009) Learning bounded treewidth Bayesian networks. In: Koller D, Schuurmans D, Bengio Y, Bottou L (eds) Advances in neural information processing systems, vol 21. Curran Associates, Red Hook, pp 417–424

    Google Scholar 

  13. Fariselli P, Riccobelli P, Casadio R (1999) Role of evolutionary information in predicting the disulfide-bonding state of cysteine in proteins. Proteins 36(3):340–346

    Article  Google Scholar 

  14. Frasconi P, Passerini A, Vullo A (2002) A two-stage SVM architecture for predicting the disulfide bonding state of cysteines. In: Proceedings of the IEEE workshop on neural networks for signal processing, pp 25–34

  15. Freno A, Trentin E (2011) Hybrid random fields: a scalable approach to structure and parameter learning in probabilistic graphical models. Springer, New York

    Book  Google Scholar 

  16. Freno A, Trentin E, Gori M (2009) A hybrid random field model for scalable statistical learning. Neural Netw 22:603–613

    Article  Google Scholar 

  17. Freno A, Trentin E, Gori M (2009) Scalable pseudo-likelihood estimation in hybrid random fields. In: Elder J, Fogelman-Souli F, Flach P, Zaki M (eds) Proceedings of the 15th ACM SIGKDD conference on knowledge discovery and data mining (KDD 2009), pp 319–327. ACM

  18. Freno A, Trentin E, Gori M (2009) Scalable statistical learning: a modular Bayesian/Markov network approach. In: Proceedings of the international joint conference on neural networks (IJCNN 2009), pp 890–897. IEEE

  19. Freno A, Trentin E, Gori M (2010) Kernel-based hybrid random fields for nonparametric density estimation. In: Proceedings of the 19th European conference on artificial intelligence (ECAI), pp 427–432. IOS Press, Amsterdam, The Netherlands

  20. Friedman N (1998) The Bayesian structural EM algorithm. In: Proceedings of the fourteenth conference on uncertainty in artificial intelligence, pp 129–138. Morgan Kaufmann, San Francisco, USA

  21. Ghahramani Z (1998) Learning dynamic Bayesian networks. In: Adaptive processing of sequences and data structures, pp 168–197. Springer

  22. Gilks WR, Richardson S, Spiegelhalter D (1996) Markov chain Monte Carlo in practice. CRC Press, Boca Raton

    MATH  Google Scholar 

  23. Grünwald PD (2007) The minimum description length principle. The MIT Press, Cambridge

    Google Scholar 

  24. Hertz J, Krogh A, Palmer RG (1991) Introduction to the theory of neural computation. Addison Wesley, Redwood

    Google Scholar 

  25. Kindermann R, Snell JL (1980) Markov random fields and their applications. American Mathematical Society, Providence

    Book  Google Scholar 

  26. Koller D, Friedman N (2009) Probabilistic graphical models: principles and techniques. MIT Press, Cambridge

    MATH  Google Scholar 

  27. Lafferty J, McCallum A, Pereira F (2001) Conditional random fields: probabilistic models for segmenting and labeling sequence data. In: Proceedings of 18th international conference on machine learning, pp 282–289. Morgan Kaufmann, San Francisco, CA

  28. Lauritzen SL (1996) Graphical models. Oxford University Press, Oxford

    MATH  Google Scholar 

  29. Lin C, Yang C, Hor C, Huang K (2010) Disulfide bonding state prediction with SVM based on protein types. In: Proceedings of the fifth IEEE international conference on bio-inspired computing: theories and applications, pp 1436–1442

  30. Mandic DP, Chambers J (2001) Recurrent neural networks for prediction: learning algorithms, architectures and stability. Wiley, New York

    Book  Google Scholar 

  31. Mitchell TM (1997) Machine learning. McGraw-Hill, New York

    MATH  Google Scholar 

  32. Mozer MC (1989) A focused backpropagation algorithm for temporal pattern recognition. Complex Syst 3(1):349–381

    MathSciNet  MATH  Google Scholar 

  33. Neal R, Hinton GE (1999) A view of the EM algorithm that justifies incremental, sparse, and other variants. In: Jordan MI (ed) Learning in graphical models. MIT Press, Cambridge, pp 355–368

    Google Scholar 

  34. Neapolitan RE (2004) Learning Bayesian networks. Prentice Hall, Upper Saddle River

    Google Scholar 

  35. Pearl J (1985) Bayesian networks: a model of self-activated memory for evidential reasoning. In: Proceedings of the 7th conference of the cognitive science society, University of California, Irvine, pp 329–334

  36. Pearl J (1988) Probabilistic reasoning in intelligent systems. Morgan Kaufmann, San Francisco

    MATH  Google Scholar 

  37. Rabiner LR (1989) A tutorial on hidden Markov models and selected applications in speech recognition. In: Proceedings of the IEEE, vol 77, no 2, pp 257–286

    Article  Google Scholar 

  38. Rust R, Schmittlein D (1985) A Bayesian cross-validated likelihood method for comparing alternative specifications of quantitative models. Mark Sci 4(1):20–40

    Article  Google Scholar 

  39. Savojardo C, Fariselli P, Alhamdoosh M, Martelli PL, Pierleoni A, Casadio R (2011) Improving the prediction of disulfide bonds in eukaryotes with machine learning methods and protein subcellular localization. Bioinformatics 27(16):2224–2230

    Article  Google Scholar 

  40. Scanagatta M, de Campos CP, Corani G, Zaffalon M (2015) Learning Bayesian networks with thousands of variables. In: Cortes C, Lawrence ND, Lee DD, Sugiyama M, Garnett R (eds) Advances in neural information processing systems, vol 28. Curran Associates, Red Hook, pp 1864–1872

    Google Scholar 

  41. Scanagatta M, Corani G, de Campos CP, Zaffalon M (2016) Learning treewidth-bounded Bayesian networks with thousands of variables. In: Lee DD, Sugiyama M, Luxburg UV, Guyon I, Garnett R (eds) Advances in neural information processing systems, vol 29. Curran Associates, Red Hook, pp 1462–1470

    Google Scholar 

  42. Shoombuatong W, Traisathit P, Prasitwattanaseree S, Tayapiwatana C, Cutler RW, Chaijaruwanich J (2011) Prediction of the disulphide bonding state of cysteines in proteins using conditional random fields. IJDMB 5(4):449–464

    Article  Google Scholar 

  43. Singh R (2008) A review of algorithmic techniques for disulfide-bond determination. Brief Funct Genomic Proteomic 7(2):157–172

    Article  Google Scholar 

  44. Trentin E (2015) Maximum-likelihood normalization of features increases the robustness of neural-based spoken human–computer interaction. Pattern Recogn Lett 66(C):71–80

    Article  Google Scholar 

  45. Trentin E, Bongini M (2012) Towards a novel probabilistic graphical model of sequential data: fundamental notions and a solution to the problem of parameter learning. In: Proceedings of the 5th INNS IAPR TC 3 GIRPR workshop on artificial neural networks in pattern recognition, pp 72–81. Springer

  46. Trentin E, Cattoni R (1999) Learning perception for indoor robot navigation with a hybrid hidden Markov model/recurrent neural networks approach. Connect Sci 11(3–4):243–265

    Article  Google Scholar 

  47. Trentin E, Giuliani D (2001) A mixture of recurrent neural networks for speaker normalisation. Neural Comput Appl 10(2):120–135

    Article  Google Scholar 

  48. Witten IH, Frank E (2005) Data mining, 2nd edn. Morgan Kaufmann, San Francisco

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Edmondo Trentin.

Appendices

A Formal Definition of HRF and Related Notions

The material presented in this Appendix and in the next one is drawn from [15]. HRFs are aimed at representing joint probability distributions underlying sets of random variables. The definition of HRF involves a few preliminary notions concerning directed and undirected graphs. First, let us define the concept of (directed) union of directed graphs:

Definition 1

If \(\mathcal {G}_{1} = (\mathcal {V}_{1}, \mathcal {E}_{1})\) and \(\mathcal {G}_{2} = (\mathcal {V}_{2}, \mathcal {E}_{2})\) are directed graphs, then the (directed) union of \(\mathcal {G}_{1}\) and \(\mathcal {G}_{2}\) (denoted by \(\mathcal {G}_{1} \cup \mathcal {G}_{2}\)) is the directed graph \(\mathcal {G} = (\mathcal {V}, \mathcal {E})\) such that \(\mathcal {V} = \mathcal {V}_{1} \cup \mathcal {V}_{2}\) and \(\mathcal {E} = \mathcal {E}_{1} \cup \mathcal {E}_{2}\).

Given a set of directed graphs \(\mathcal {G}_{1}, \ldots , \mathcal {G}_{n}\), the directed union \(\mathcal {G} = \bigcup _{i=1}^{n} \mathcal {G}_{i}\) results from iterated application of the binary union operator. Similarly, we define the notion of (undirected) union for undirected graphs as follows:

Definition 2

If \(\mathcal {G}_{1} = (\mathcal {V}_{1}, \mathcal {E}_{1})\) and \(\mathcal {G}_{2} = (\mathcal {V}_{2}, \mathcal {E}_{2})\) are undirected graphs, then the (undirected) union of \(\mathcal {G}_{1}\) and \(\mathcal {G}_{2}\) (denoted by \(\mathcal {G}_{1} \cup \mathcal {G}_{2}\)) is the undirected graph \(\mathcal {G} = (\mathcal {V}, \mathcal {E})\) such that \(\mathcal {V} = \mathcal {V}_{1} \cup \mathcal {V}_{2}\) and \(\mathcal {E} = \mathcal {E}_{1} \cup \mathcal {E}_{2}\).

Now, if \(\mathcal {G} = (\mathcal {V}, \mathcal {E})\) is a directed graph, then we say that \(\mathcal {G}^{*} = (\mathcal {V}^{*}, \mathcal {E}^{*})\) is the undirected version of \(\mathcal {G}\) if \(\mathcal {V}^{*} = \mathcal {V}\) and \(\mathcal {E}^{*} = \{\{X_{i}, X_{j}\}: (X_{i}, X_{j}) \in \mathcal {E}\}\). Thus, we also define the undirected union for a pair of directed graphs:

Definition 3

If \(\mathcal {G}_{1} = (\mathcal {V}_{1}, \mathcal {E}_{1})\) and \(\mathcal {G}_{2} = (\mathcal {V}_{2}, \mathcal {E}_{2})\) are directed graphs, then the undirected union of \(\mathcal {G}_{1}\) and \(\mathcal {G}_{2}\) (denoted by \(\mathcal {G}_{1} \Cup \mathcal {G}_{2}\)) is the (undirected) graph \(\mathcal {G}^{*} = \mathcal {G}_{1}^{*} \cup \mathcal {G}_{2}^{*}\) such that \(\mathcal {G}_{1}^{*}\) and \(\mathcal {G}_{2}^{*}\) are the undirected versions of \(\mathcal {G}_{1}\) and \(\mathcal {G}_{2}\) respectively.

Finally, a HRF can be defined as follows:

Definition 4

Let \(\mathbf {X}\) be a set of random variables \(X_{1}, \ldots , X_{n}\). A hybrid random field for \(X_{1}, \ldots , X_{n}\) is a set of Bayesian networks \(BN_{1}, \ldots , BN_{n}\) (with DAGs \(\mathcal {G}_{1}, \ldots , \mathcal {G}_{n}\)) such that:

  1. 1.

    Each \(BN_{i}\) contains \(X_{i}\) plus a subset \(\mathcal {R}(X_{i})\) of \(\mathbf {X} \setminus \{X_{i}\}\);

  2. 2.

    If \(\mathcal {MB}_{i}(X_{i})\) denotes the Markov blanket [36] of \(X_{i}\) in \(BN_{i}\) (i.e. the set of the parents, the children, and the parents of the children of \(X_{i}\) in \(\mathcal {G}_{i}\)) and \(P(X_{i} \mid \mathcal {MB}_{i}(X_{i}))\) is the conditional distribution of \(X_{i}\) given \(\mathcal {MB}_{i}(X_{i})\), as derivable from \(BN_{i}\), then at least one of the following conditions is satisfied:

    1. (a)

      The directed graph \(\mathcal {G} = \bigcup _{i=1}^{n} \mathcal {G}_{i}\) is acyclic and there is a Bayesian network \(h_{\mathcal {G}}\) with DAG \(\mathcal {G}\) such that, for each \(X_{i}\) in \(\mathbf {X}\), \(P(X_{i} \mid \mathcal {MB}_{i}(X_{i})) = P(X_{i} \mid \mathcal {MB}_{\mathcal {G}}(X_{i}))\), where \(\mathcal {MB}_{\mathcal {G}}(X_{i})\) is the Markov blanket of \(X_{i}\) in \(h_{\mathcal {G}}\) and \(P(X_{i} \mid \mathcal {MB}_{\mathcal {G}}(X_{i}))\) is the conditional distribution of \(X_{i}\) given \(\mathcal {MB}_{\mathcal {G}}(X_{i})\), as entailed by \(h_{\mathcal {G}}\);

    2. (b)

      There is a Markov random field \(h_{\mathcal {G}^{*}}\) with graph \(\mathcal {G}^{*}\) such that \(\mathcal {G}^{*} = \mathcal {G}_{1} \Cup \ldots \Cup \mathcal {G}_{n}\) and, for each \(X_{i}\) in \(\mathbf {X}\), \(P(X_{i} \mid \mathcal {MB}_{i}(X_{i})) = P(X_{i} \mid \mathcal {MB}_{\mathcal {G}^{*}}(X_{i}))\), where \(\mathcal {MB}_{\mathcal {G}^{*}}(X_{i})\) is the Markov blanket of \(X_{i}\) in \(h_{\mathcal {G}^{*}}\) and \(P(X_{i} \mid \mathcal {MB}_{\mathcal {G}^{*}}(X_{i}))\) is the conditional distribution of \(X_{i}\) given \(\mathcal {MB}_{\mathcal {G}^{*}}(X_{i})\), as derived from \(h_{\mathcal {G}^{*}}\).

The elements of \(\mathcal {R}(X_{i})\) are called ‘relatives of \(X_{i}\)’. That is, the relatives of a node \(X_{i}\) in a HRF are the nodes appearing in graph \(\mathcal {G}_{i}\) (except for \(X_{i}\) itself). An example of HRF over the variables \(X_1, \ldots , X_4\) is shown in Fig. 10.

Fig. 10
figure 10

The graphical components of a hybrid random field for the variables \(X_{1}, \ldots , X_{4}\). Since each node \(X_{i}\) has its own Bayesian network (where nodes in \(\mathcal {MB}_{i}(X_{i})\) are shaded), there are four different DAGs \({\mathcal {G}}_1, \ldots , {\mathcal {G}}_4\). Relatives of \(X_{i}\) that are not in \(\mathcal {MB}_{i}(X_{i})\) are dashed

Two fundamental properties of HRFs stem from the following Modularity Theorem, whose proof is given in [15].

Theorem 1

Suppose that h is a hybrid random field for the random vector \(\mathbf {X} = (X_{1}, \ldots , X_{n})\), and let h be made up by Bayesian networks \(BN_{1}, \ldots , BN_{n}\) (with DAGs \(\mathcal {G}_{1}, \ldots , \mathcal {G}_{n}\)). Then, if each conditional distribution \(P(X_{i} \mid \mathcal {MB}_{i}(X_{i}))\) is strictly positive (where \(1 \le i \le n\)), h has the following properties:

  1. 1.

    An ordered Gibbs sampler [22] applied to h, i.e. an ordered Gibbs sampler which is applied to the conditional distributions \(P(X_{1} \mid \mathcal {MB}_{1}(X_{1})), \ldots , P(X_{n} \mid \mathcal {MB}_{n}(X_{n}))\), defines a joint probability distribution over \(\mathbf {X}\) via its (unique) stationary distribution \(P(\mathbf {X})\);

  2. 2.

    For each variable \(X_{i}\) in \(\mathbf {X}\), \(P(\mathbf {X})\) is such that \(P(X_{i} \mid \mathbf {X} \setminus \{X_{i}\}) = P(X_{i} \mid \mathcal {MB}_{i}(X_{i}))\).

The first condition in Theorem 1 is exploited in D-HRFs in order to express the overall likelihood of the D-HRF over an input sequence in terms of local, state-specific joint probability distributions (namely, the emission probabilities) modeled via HRFs. In turn, we refer to the second condition in Theorem 1 as the modularity property. Based on the modularity property, the set \(\mathcal {MB}_{i}(X_{i})\) is a Markov blanket of \(X_{i}\) in \(\mathbf {X}\).

B Likelihood Versus Pseudo-Likelihood in HRFs

Let \({{\mathcal {X}}} = \{X_{1}, \ldots , X_{n}\}\) be any set of n discrete random variables of interest. In order to extract the joint probability \(P({{\mathcal {X}}})\) from a HRF [15], Gibbs sampling techniques need to be used [15]. Unfortunately, Gibbs sampling can be computationally intractable [22]. Therefore, in practice we can follow in the footsteps of [3] and resort to the pseudo-likelihood function, defined as :

$$\begin{aligned} P^{*}({{\mathcal {X}}}) = \prod _{i=1}^{n} P(X_{i} \mid {{\mathcal {X}}} \setminus \{X_{i}\}) \end{aligned}$$
(29)

A discussion of the theoretical properties of the pseudo-likelihood is offered by [26]. Due to the modularity property, Eq. (29) can be rewritten as:

$$\begin{aligned} P^{*}({{\mathcal {X}}}) = \prod _{i=1}^{n} P(X_{i} \mid \mathcal {MB}_{i}(X_{i})) \end{aligned}$$
(30)

Therefore, in order to measure the pseudo-likelihood \(P^{*}({{\mathcal {X}}})\) of the HRF given \({{\mathcal {X}}}\) we only need to be able to compute the conditional distribution of each node \(X_{i}\) given the state \(mb_{i}(X_{i})\) of \(\mathcal {MB}_{i}(X_{i})\). A discussion of how this can be done in a simple and efficient way is given in [15].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bongini, M., Freno, A., Laveglia, V. et al. Dynamic Hybrid Random Fields for the Probabilistic Graphical Modeling of Sequential Data: Definitions, Algorithms, and an Application to Bioinformatics. Neural Process Lett 48, 733–768 (2018). https://doi.org/10.1007/s11063-017-9730-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11063-017-9730-3

Keywords

Navigation