Skip to main content
Log in

An ensemble approach to the evolution of complex systems

  • Published:
Journal of Biosciences Aims and scope Submit manuscript

Abstract

Adaptive systems frequently incorporate complex structures which can arise spontaneously and which may be non-adaptive in the evolutionary sense. We give examples from phase transition and fractal growth to develop the themes of cooperative phenomena and pattern formation. We discuss RNA interference and transcriptional gene regulation networks, where a major part of the topological properties can be accounted for by mere combinatorics. A discussion of ensemble approaches to biological systems and measures of complexity is presented, and a connection is established between complexity and fitness.

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.

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6

Similar content being viewed by others

References

  • Andrieux D and Gaspard P 2008 Nonequilibrium generation of information in copolymerization process. Proc. Natl. Acad. Sci. USA 105 9516–9521

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Albert R and Barabási A-L 2002 Statistical mechanics of complex networks. Rev. Mod. Phys. 74 47–97

    Google Scholar 

  • Alberts B, Johnson A, Lewis J, Raff M, Roberts K and Walter P 2002 Molecular biology of the cell (New York: Garland Science - Taylor and Francis)

    Google Scholar 

  • Almeida R and Allshire RC 2005 RNA silencing and genome regulation. Trends Cell Biol. 15 251–258

    Article  CAS  PubMed  Google Scholar 

  • Almirantis Y and Provata A 1999 Scaling properties of coding and non-coding DNA sequences. J. Stat. Phys. 97 233–262

    Google Scholar 

  • Arpağ G 2011 A stochastic model for microRNA – messenger RNA Networks. Graduation Thesis (Undergraduate Degree), Department of Physics, Istanbul Technical University, Istanbul, Turkey

  • Atlan H 1987 Self-creation of meaning. Phys. Script. 36 563–576

  • Atlan H and Cohen IR 1998 Immune information, self-organization and meaning. Intl. Immun. 10 711–717

    Google Scholar 

  • Balcan D and Erzan A 2004 Random model for RNA interference yields scale free network. Eur. Phys. J. B 38 253–260

    Article  CAS  Google Scholar 

  • Balcan D and Erzan A 2007 Content-based networks: a pedagogical overview. Chaos 17 026108

    Google Scholar 

  • Balcan D, Kabakçıoğlu A, Mungan M and Erzan A 2007 The information coded in the yeast response elements accounts for most of the topological properties of its transcriptional regulation network. PLoS ONE 2 e501

    Article  PubMed Central  PubMed  Google Scholar 

  • Barabási AL and Oltvai ZN 2004 Network biology: understanding the cell’s functional organization. Nat. Rev. Genet. 5 101–113

    Google Scholar 

  • Basu M, Bhattacharyya NP, Mohanty PK 2014 Link-weight distribution of microRNA co-target networks exhibit universality. EPL 105 28007

    Google Scholar 

  • Ben-Jacob E, Cohen I, Czirók A, Vicsek T and Gutnick DL 1997 Chemomodulation of cellular movement, collective formation of vortices by swarming bacteria, and colonial development. Physica A: Stat. Mech. Appl. 238 181–197

    Article  Google Scholar 

  • Bollobás B 2001 Random graphs (Cambridge UK: Cambridge University Press)

  • Buldyrev SV, Dokholyan NV, Goldberger AL, Havlin S, Peng C-K, Stanley HE and Viswanathan GM 1998 Analysis of DNA sequences using methods of statistical physics. Physica A 249 430–438

    Article  CAS  Google Scholar 

  • Carthew RW and Southeimer EJ 2009 Origins and mechanisms of miRNAs and siRNAs. Cell 136 642–655

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Ciliberti S, Martin OC and Wagner A 2007 Robustness can evolve gradually in complex regulatory gene networks with varyingtopology. PLoS Comput. Biol. 3 e15

  • Crooks GE 1999 Entropy production fluctuation theorem and the non-equilibrium work relation for free energy differences. Phys. Rev. E 60 2721–2726

    Article  CAS  Google Scholar 

  • Crooks GE 2000 Path-ensemble averages in systems driven far from equilibrium. Phys. Rev. E 61 2361–2366

    Article  CAS  Google Scholar 

  • Chen K and Rajewski N 2007 The evolution of gene regulation by transcription factors and microRNAs. Nat. Genet. 8 93–103

    Article  CAS  Google Scholar 

  • Dokholyan NV, Buldyrev SV, Havlin S, and Stanley HE 1997 Distribution of base pair repeats in coding and noncoding DNA sequences. Phys. Rev. Lett. 79 5182–5185

    Google Scholar 

  • Erdös P and Renyi A 1959 Publ. Mat. (Debrecen) 6 290ff. Erdös P and Renyi A 1960 Publ. Mat. Inst . Hung. Acad . Sci. 5 17 and Erdös P and Renyi A 1961 Bull. Inst . Int. Stat . 38 343; cited in Albert R and Barabási A-L 2002 Statistical mechanics of complex networks. Rev. Mod. Phys. 74 47–97

  • Erzan A, Pietronero L and Vespignani A 1995 The fixed scale transformation approach to fractal growth. Rev. Mod. Phys. 67 545–604

    Google Scholar 

  • Feschotte C and Gilbert C 2012 Endogenous viruses: insights into viral evolution and impact on host biology. Nat. Rev. Genet. 13 283–296

    Google Scholar 

  • Fisher D, Lässig M and Shraiman B 2013 Evolutionary dynamics and statistical physics. J. Stat. Mech. doi:10.1088/1742-5468/2013/01/N01001

    Google Scholar 

  • Fisher RA 1930 The genetical theory of natural selection (Oxford, UK: Oxford University Press); reproduced in 2000 The genetical theory of natural selection: A complete varorium edition (ed) H Bennet (Oxford, UK: Oxford University Press)

  • Frisch U 1995 Turbulence (Cambridge: Cambridge University Press)

    Google Scholar 

  • Fujikawa H and Matsushita M 1989 Fractal growth of Bacillus subtilis on agar plates. J. Phys. Soc. Jpn. 58 3875–3878

    Google Scholar 

  • Gould S and Lewontin RC 1979 The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist program. Proc. R. Soc. Lond. B 205 581–598

    Article  CAS  PubMed  Google Scholar 

  • Grassberger P 1986a How to measure self-generated complexity. Physica A 140 319–325

    Article  Google Scholar 

  • Grassberger P 1986b Toward a quantitative theory of self-generated complexity. Int. J. Theo. Phys. 25 907–938

    Article  Google Scholar 

  • Grassberger P 2012 Randomness, information, and complexity. arxiv:1208.3459v1

  • Griffiths-Jones S, Saini, HK, van Dongen S and Enright AJ 2008 miRBase: Tools for microRNA genomics. Nucleic Acids Res. 36 D154–D158

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Grillo G, Turi T, Licciulli F, Mignone F, Liuni S, Banfi S, Gennarino VA, Horner DA, Pavesi G, Picardi E and Pesole G 2010 UTRdb and UTRsite (RELEASE 2010): a collection of sequences and regulatory motifs of the untranslated regions of eukaryotic mRNAs. Nucleic Acids Res. 38 D75–D80

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Grimson A, Kai-How Farh K, Johnston WK, Garrett-Engele P, Lim LP and Bartel DP 2007 MicroRNA targeting specificity in mammals: Determinants beyond seed pairing. Mol. Cell 27 91–105

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Grosse I, Herzel H, Buldyrev SV and Stanley HE 2000 Species independence of mutual information in coding and non-coding DNA. Phys. Rev. E 61 5624–2629

    Article  CAS  Google Scholar 

  • Hannon GJ 2002 RNA interference. Nature 418 244–251

    Article  CAS  PubMed  Google Scholar 

  • Hao B-L, Lee HC and Shu-Yu Z 2000 Fractals related to long DNA sequences and complete genomes. Chaos Sol. Fractals 11 825–836

    Google Scholar 

  • Harbison CT, Gordon DB, Lee TI, Rinaldi NJ, Macisaac KD, et al. 2004 Transcriptional regulatory code of a eukaryotic genome. Nature 431 99–104

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Hofacker I 2007 How microRNAs choose their targets. Nat. Genet. 39 1191–1192

    Google Scholar 

  • Houchmandzade B 2014 The remarkable discreteness of being. J. Biosci. 39 249–258

    Google Scholar 

  • Jarzynski C 1997 Nonequilibrium equality for free energy differences. Phys. Rev. Lett. 78 2690–2693

    Google Scholar 

  • Jarzynski C 2008 Thermodynamics of writing a random polymer. Proc. Natl. Acad. Sci. 105 9451–9452

    Google Scholar 

  • Kapusta A, Kronenberg Z, Lynch VJ, Zhao X, Ramsey LA, Borque G, Yandell M and Feschotte C 2013 Transposable elements are major contributors to the origin, diversification and regulation of vertebrate long non-coding RNAs. PLoS Genet. 9 31003470

  • Keskin O, Gursoy A, Ma B and Nussinov R 2008 Principles of protein − protein interactions: what are the preferred ways for proteins to interact? Chem. Rev. 108 1225−1244

  • Kondepudi D and Prigogine I 1998 Modern thermodynamics: From heat engines to dissipative structures (New York: Wiley)

    Google Scholar 

  • Kolmogorov AN 1965 Three approaches to the quantitative definition of information. Problems Inf. Trans. 1 1–7

    Google Scholar 

  • Kolmogorov AN 1983 Combinatorial foundation of information theory and the calculus of probabilities. Russ. Math. Surveys 38 27–36

    Article  Google Scholar 

  • Lamberti PW, Martin MT, Plastino A and Rosso OA 2004 Intensive entropic non-triviality measure. Physica A 334 119–131

    Article  Google Scholar 

  • Lopez-Ruiz R, Mancini HL and Calbet X 1995 A statistical measure of complexity. Phys. Lett. A 209 321–326

    Article  CAS  Google Scholar 

  • Lynch M 2007 The evolution of genetic networks by non-adaptive processes. Nat. Rev. Genet. 8 803–813

    Google Scholar 

  • Mahajan VS, Drake A and Chen J 2008 Virus-specific host miRNAs. Trends Immun. 30 1–7

    Article  Google Scholar 

  • Maini PK 2004 Using mathematical models to help understand biological pattern formation. Comptes Rendu Biologies 327 225–234

    Article  Google Scholar 

  • Majoros WH and Ohler U 2007 Spatial preferences of microRNA targets in 3' untranslated regions. BMC Genomics 8 doi:10.1186/1471-2164-8-152

  • Malkoç B, Balcan D and Erzan A 2011 Information content based model for the topological properties of the gene regulatory network of E. coli. J. Theor. Biol. 203 281–294

    Google Scholar 

  • Mandelbrot BB 1982 The fractal geometry of nature (New York: WH Freeman and Co.)

  • Mantegna RN, Buldyrev SV, Goldberger AL, Havlin S, Peng C-K, Simons M and Stanley HE 1995 Systematic analysis of coding and non-coding DNA sequences using statistical linguistics. Phys. Rev. E 52 2939–2950

    Article  CAS  Google Scholar 

  • Martin MT, Plastino A and Rosso OA 2003 Statistical complexity and disequilibrium. Phys. Lett. A 311 126–132

    Article  CAS  Google Scholar 

  • Matsushita M and Fujikawa H 1990 Diffusion-limited growth in bacterial colony formation. Physica A 168 498–506

    Article  CAS  Google Scholar 

  • Matsushita M, Hiramatsu F, Kobayashi N, Ozawa T, Yamazaki Y and Matsuyama T 2004 Colony formation in bacteria: experiments and modeling. Biofilms 1 305–317

    Article  Google Scholar 

  • Mello CC and Conte D 2004 Revealing the world of RNA interference. Nature 431 338–342

    Article  CAS  PubMed  Google Scholar 

  • Mello CC and Fire A 2000 Functional anatomy of a dsRNA trigger: Differential requirement for the two trigger strands in RNA interference. Mol. Cell 6 1077–1087

    Google Scholar 

  • Mookherjee S, Sinha M, Mukhopadhyay S, Bhattacharyya N and Mohanty PK 2009 Analysis of clustered microRNA in biological pathways. Online J. Bioinformatics 10 280–296

    Google Scholar 

  • Morange M 2008 What history tells us XIV. Regulation of gene expression by non-coding RNAs: the early steps. J. Biosci. 33 327–331

    Article  CAS  PubMed  Google Scholar 

  • Morange M 2012 What history tells us XXIX. Transfers from plant biology: From cross protection to RNA interference and DNA vaccination. J. Biosci. 37 949–952

    Article  PubMed  Google Scholar 

  • Mourier T and Willerslev E 2009 Retrotransposons and non–protein coding RNAs. Brief. Funct. Genom. Proteom. doi:10.1093/bfgp/elp036

    Google Scholar 

  • Mungan M, Kabakçıoğlu A, Balcan D and Erzan A 2005 Analytical solution of a stochastic model for gene regulation. J. Phys. A 38 9599–9620

    Article  Google Scholar 

  • Mustonen V and Lässig M 2010 Fitness flux and ubiquity of adaptive evolution. Proc. Natl. Acad. Sci. USA 107 4248–4253

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Newman SA and Forgacs G 2005 Complexity and self-organization in biological development; in Complexity in chemistry, biology and ecology (eds) DD Bonchev and D Rouvray (Heidelberg: Springer)

    Google Scholar 

  • Nicolis G and Prigogine I 1977 Self-organization in nonequilibrium systems: From dissipative structures to order through fluctuations (New York: Wiley)

    Google Scholar 

  • Niemeyer L, Pietronero L and Wiesmann HJ 1984 Fractal dimension of dielectric breakdown. Phys. Rev. Lett. 52 1033–1036

    Google Scholar 

  • Novina CD and Sharp PA 2004 The RNAi revolution. Nature 430 161–164

    Article  CAS  PubMed  Google Scholar 

  • Orr HA 2000 Adaptation and the cost of complexity. Evolution 54 13–20

    Article  CAS  PubMed  Google Scholar 

  • Peng C-K, Buldyrev SV, Goldberger AL, Havlin S, Sciortino F, Simons M and Stanley HE 1992 Long range correlations in nucleotide sequences. Nature 356 168–170

    Article  CAS  PubMed  Google Scholar 

  • Peng C-K, Havlin S, Simons M, Stanley HE and Goldberger AL, 1994 Mosaic organization of DNA nucleotides. Phys. Rev. E 49 1685–1689

    Google Scholar 

  • Pietronero L, Erzan A and Evertsz C 1988 Theory of fractal growth. Phys. Rev. Lett. 61 861–864

    Google Scholar 

  • Piriyapongsa J and Jordan IK 2007 A family of human microRNA genes from miniature inverted-repeat transposable elements. PLoS ONE 2 e203

    Article  PubMed Central  PubMed  Google Scholar 

  • Román-Roldán R, Bernaola-Galván P and Oliver JL 1998 Sequence compositional complexity of DNA. Phys. Rev. Lett. 80 1344–1347

    Google Scholar 

  • Rosso OA, Craig H and Moscato P 2009 Shakespeare and other English Renaissance authors as characterized by information theory complexity quantifiers. Physica A 388 916–926

    Article  Google Scholar 

  • Salazar-Ciudad I, Newman SA and Sole RV 2001 Phenotypic and dynamical transitions in model genetic networks I. Emergence of patterns and genotype-phenotype relationships. Evol. Dev. 3 84–94

    Article  CAS  PubMed  Google Scholar 

  • Sadhu T and Dhar D 2010 Pattern formation in growing sandpiles with multiple sources or sinks. J. Stat. Phys. 138 815–837

    Google Scholar 

  • Shannon CE 1948 A mathematical theory of communication. Bell Syst. Tech. J. 27 379–423 and 623–656

  • Stanley HE 1971 Introduction to phase transitions and critical phenomena (New York: Oxford University Press)

    Google Scholar 

  • Stanley HE, Buldyrev SV, Goldberger AL, Havlin S, Peng C-K and Simons M 1999 Scaling features of non-coding DNA. Physica A 273 1–18

    Article  CAS  PubMed  Google Scholar 

  • Stauffer D and Aharony A 1992 Introduction to percolation theory (London: Taylor and Francis)

    Google Scholar 

  • Teixeira MC, Monteiro P, Jain P, Tenreiro S, Fernandes AR, et al. 2006 The YEASTRACT database: a tool for the analysis of transcription regulatory associations in Saccharomyces cerevisiae. Nucleic Acids Res. 34 D446–D451

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Thomas DW 1942 On growth and form (New York: Dover Publications – 1992 edition)

  • Tong AHY, Lesage G, Bader GD, Ding H, Xu H, et al. 2004 Global mapping of the yeast genetic interaction network. Science 303 808–813

    Article  CAS  PubMed  Google Scholar 

  • Turing AM 1952 The chemical basis of morphogenesis. Phil. Trans. R. Soc. Lond. B 237 37–72

    Article  Google Scholar 

  • Turing AM 1936 On computable numbers, with an application to the Entscheidungs problem. Prof. Lond. Math. Soc. 42 230–265

    Google Scholar 

  • Veksler-Lublinsky I, Shemer-Avni Y, Kedem K, Ziv-Ukelson 2010 Gene bi-targeting by viral and human miRNAs. BMC Bioinformatics 11 249

    Article  PubMed Central  PubMed  Google Scholar 

  • Wagner A 1994 Evolution of gene networks by gene duplications: a mathematical model and its implications on genome organization. Proc. Natl. Acad. Sci. USA 91 4387–4391

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Witten TA and Sander LM 1983 Diffusion limited aggregation. Phys. Rev. B 27 5686–5697

    Article  Google Scholar 

  • Wolfram S 1983 Statistical mechanics of cellular automata. Rev. Mod. Phys. 55 601–644

    Google Scholar 

  • Zhuo Y, Gao G, Shi J, Zhou X and Wang X 2013 miRNAs: Biogenesis, origin and evolution, functions on virus-host interaction. Cell Physiol. Biochem. 32 499–510

    Google Scholar 

  • Zurek WH 1989 Thermodynamic cost of computation, algorithmic complexity and the information metric. Nature 341 119–124

    Article  Google Scholar 

Download references

Acknowledgments

AE would like to thank the organizers of the Almora meeting on ‘Individuals and Groups’ for a very challenging and stimulating conference. The work on RNAi modelling is in partial fulfillment of the requirements for Göker Arpağ’s undergraduate degree. We would like to express our gratitude to Dr Mahashweta Basu for providing us with the curated version (double counting of some miRNAs were removed) of the predicted ( http://www.ebi.ac.uk/enright-srv/microcosm/cgi-bin/targets/v5/download.pl ) miRNA-mRNA network for the four animals. The work on modelling the transcriptional gene regulatory network of yeast went towards the partial fulfillment of the PhD degree requirements of Duygu Balcan. We would like to thank Drs. Renee Borges and Tejas Murthy, Indian Institute of Science, Bangalore, for their contribution to the discussion on measures of complexity. It is a pleasure to acknowledge many useful interactions with Duygu Balcan, Muhittin Mungan, Alkan Kabakcioglu and Pradeep K Mohanty. AE’s travel to Almora was made possible with a grant from the Istanbul Technical University BAPSO Scientific Research Fund.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ayşe Erzan.

Additional information

[Arpağ G and Erzan A 2014 An ensemble approach to the evolution of complex systems. J. Biosci. 39 1–22] DOI 10.1007/s12038-013-9394-8

Supplementary materials pertaining to this article are available on the Journal of Biosciences Website at http://www.ias.ac.in/jbiosci/april2014/supp/Arpag.pdf

Electronic supplementary material

Below is the link to the electronic supplementary material.

ESM 1

(PDF 349 kb)

Appendices

Appendix 1 Ensembles, entropy, information

The terminology relating to probabilities and probability distributions shows a great deal of variance (!) between neighbouring disciplines, and this is why we found it necessary to supply this appendix. As physicists, we may be using a slightly different convention than the intended readership of this article.

1.1 Appendix 1.1 Ensembles and probabilities

In order to talk about the probability of occurrence of a certain configuration, event or value, one should first specify what class of systems one has in mind, where the frequency of a configuration, an event or the particular value of a stochastic variable, are to be sought. This class of systems is called the statistical ensemble, or ensemble for short. Theoretically, this imaginary ensemble consists of a very large number of copies of the system under consideration, prepared in the same way and subject to the same external constraints.

The ‘copy of a given system’ may be as simple as the casting of a die, or as complicated as a particular realization of a model network for the gene regulatory system of E. coli. The number of occurrences (N x ) of a given value or event, in this particular ensemble (say the number X on the up face of the die, or the number of genes having X interactions with other genes) divided by the size of the ensemble (the number of throws of the die, the total number of genes in the network) tends to the probability p(X) of (sometimes called the marginal probability, or the marginal) of this value, event, etc., as the size (N) of the ensemble is increased without bound. This relation we can express symbolically as

$$ { \lim}_{N\to \infty}\left({N}_X/N\right)=p(X). $$

The reason for taking this (ideal) limit is that in any experimental situation, either on the field, in the laboratory, or when doing a computer simulation, for any finite number of samples N, the ratio (N X /N ) itself is a fluctuating random variable. It can be rigorously shown, however, that in most cases, the variance of this ratio will tend to zero in this ideal limit, and this is how we would like to define our ‘probability distribution’, i.e. the function p(X) for all different allowed values of x.

In real-life experiments one has to do with finite populations, and therefore one has to make do with approximations to the probability distribution, in the form of p(X) ≈ (N X / N ). Often one has to average over many different runs, populations, or realizations of numerical models, to reduce the variance.

1.2 Appendix 1.2 Some statistical concepts and notation

The average, or expectation value, of a stochastic variable x taking the values x i with probabilities p i can be expressed as 〈x〉 = ∑  i p i x i . (In physics, one usually does not make a distinction between the name of the variable and the different values that it may take; instead of X in the above paragraph, here we have used x i , and summed over i.) In case the variable takes continuous values, the sum goes over to an integral; the probability distribution {p i } is then replaced by p(x)dx where p(x) is now a probability density function over allowed values of x. The statistical ensemble to which the variable x belongs is defined by the set of probabilities {p i } or the probability density function p(x).

Correlations between stochastic variables in a given ensemble are measured via the correlation function. The correlation function between two variables is a measure of the degree to which they depart from being independent of each other, usually as a function of the distance between them. Note that in case the variable x and y are independent of each other the expectation value 〈xy〉 = 〈x〉〈y〉. Thus, 〈xy〉 − 〈x〉〈y〉 ≠ 0 indicates that the variables x and y are not independent, but ‘correlated.’ The two-point correlation function

$$ C(r)=\left\langle {x}_0{x}_r\right\rangle -\left\langle {x}_0\right\rangle \left\langle {x}_r\right\rangle $$

measures the correlation between two variables x 0 and x r as a function of r, where r may indicate their distance from each other in some metric space; we expect the correlation to decrease as a function of this distance. For example, density fluctuations ρ in a fluid at neighbouring points set off by a distance r from each other would obey a correlation function which decreases exponentially with r, unless the fluid is at its critical temperature and pressure, in which case it would decrease only as a power of r.

The values that the variables x and y might take could be letters from an alphabet. Given an ensemble of letter-sequences, we could compute the two-point correlation function between characters appearing at positions along the sequence separated by r letters.

1.3 Appendix 1.3 Information content and entropy

The definition of information given by Shannon (1948) is that it is equal to the entropy for a given probability distribution p x (X) for the variable x to take on different values X,

$$ S=I=-{\displaystyle {\sum}_X{p}_x\left(\mathrm{X}\right) \log {p}_x\left(\mathrm{X}\right).} $$

Note that this definition of information is the same as the entropy up to a dimensional constant which we have set equal to unity. We will use the term information and entropy interchangeably throughout the article. We have used X to denote not necessarily one but a whole list of attributes with which a particular state may be specified, and the sum is over all such states. The probability distribution p x (X) characterizes the statistical ensemble to which the system under consideration belongs. Note that for an equiprobable distribution, I is simply the logarithm of the number of possible states. On the other hand, if any of the probabilities p x (X)

The mutual information between two probability distributions for the variables x and y, p x (X) and p y (Y) is defined as

$$ {\mathrm{I}}_M={\displaystyle \sum_{X,Y}{p}_{xy}\left(X,Y\right)}\left[ ln{p}_{xy}\left(X,Y\right)- \ln {p}_x(X)- \ln {p}_y(Y)\right] $$

where p xy (X,Y) is the joint probability distribution.

The conditional information I cond is defined by

$$ {I}_{\mathrm{cond}}={\displaystyle \sum_Y{p}_y(Y){\displaystyle \sum_X{p}_{xy}\left(X\Big|Y\right)}} \ln {p}_{xy}\left(X\Big|Y\right) $$

where p xy (X|Y) is the conditional probability distribution of x conditional upon y. Note p xy (X|Y) = p xy (X,Y)/p y (Y), which gives, I M  = I x  − I cond. We have indicated the information content of the probability distribution of x by I x , for clarity.

1.4 Appendix 1.4 Relative information and the effective length of sequences in terms of their relative information content

The probability matrix of a ‘binding site’ is found in the following way: One defines each entry of this matrix, p ij , to be the probability with which the jth site along the binding sequence for that protein is occupied by A (i = 1), T(i = 2) ,C(i = 3) or G (i = 4). Note that i runs from 1 to 4 and j runs from 1 to n, where n is the length of the consensus sequence.

The information content of a binding site is equal to the entropy of the probability matrix of the binding site,

$$ S=I=-{\displaystyle {\sum}_{i,j}{p}_{ij} \log {p}_{ij}} $$

where i is summed from 1 to 4, and j from 1 to n. The information, I, is equal to zero if no ‘mistakes’ are allowed on any of the sites, i.e. there is one unique binding sequence. On the other hand, if the binding occurs in a totally indiscriminate way, so that any letter can appear with equal probability at any of the positions along the sequence, the information (entropy) takes the largest value it can take, namely I 0 = n log 4.

Let us define the relative information to be I R = I 0I. This quantity measures how different the binding sequence is from a perfectly random one. The effective number n e of strict constraints that have to be satisfied for the binding to occur is then given by the relative entropy divided by log4; i.e. n e = I R /log 4. Note that n e will be close to zero if the binding sequence is very indiscriminate; it will be close to n if the binding is very strict.

The distance between two probability distributions P and Q may be defined as

$$ D={\displaystyle {\sum}_X{\left[P(X)-Q(X)\right]}^2.} $$

If Q is the uniform distribution, D is known as the ‘disequilibrium’ (Nicolis and Prigogine 1977), i.e. the distance of the probability distribution P from the uniform (or equiprobable) distribution.

The statistical complexity is defined as

$$ C\left[P\right]=J\left[P,Q\right]\times S(P)/{S}_{\max }{J}_{\max }, $$

where J[P,Q] is the Jensen–Shannon divergence is related to the relative information between the probability distributions P and Q. It is defined as

$$ J\left[P,Q\right]=S\left[\left(P+Q\right)/2\right]-\left[S(P)+S(Q)\right]/2. $$

The expression S [(P + Q) / 2] means, explicitly, S (U) = − Σ X U(X) log U(X), where we have set [P(X) + Q(X)] / 2 = U(X).

Given that the number of allowed states of the system is N s , the distribution Q may be chosen as the equiprobable distribution, Q(X) =1/N s , for which the entropy of the system would be maximum. Alternatively, Q can be chosen as any reference state which has a greater degree of randomness than the system at hand, for example, the probability distribution of a null model (Martin et al. 2003; Lamberti et al. 2004; Rosso et al. 2009). Note that a similar statistical complexity can be defined in terms of the ‘disequilibrium’ rather than the Jensen–Shannon divergence.

Appendix 2 Network properties

2.1 Appendix 2.1 The adjacency matrix

Any network having N nodes can be represented by an N × N matrix called the adjacency matrix. The elements of this matrix, A ij , can be assigned weights, which characterize the strength of interaction between the nodes indexed by i and j, running from 1 to N. If the network is not weighted, so that it simply has edges which are either there or absent, then the elements of the adjacency matrix are either zero or unity. If the network is directed, or if there are different types of nodes, the adjacency matrix will be asymmetric, and symmetric otherwise. In general, summing over the ith row of the matrix will give the in-degree of the ith node; similarly summing over the jth column will give the out-degree of the jth node.

$$ \begin{array}{cc}\hfill {d}_i^{\mathrm{in}}={\displaystyle \sum_j{A}_{ij}}\hfill & {d}_j^{\mathrm{out}}={\displaystyle \sum_i{A}_{ij}}\hfill \end{array} $$

For a symmetric adjacency matrix these two quantities are obviously equal to each other, and then one simply talks about the degree of a node.

2.2 Appendix 2.2 Some topological properties of networks

The ‘degree distribution’ of a network is defined as the relative frequency (in the limit of an infinite graph, the probability) of those nodes which have d other nodes connected to them by edges.

On a ‘classical’ random graph, or Erdös-Renyi (ER) network (Erdös and Renyi 1959), each pair of nodes is connected to each other with the same probability p, independently of all the other pairs. The expected degree, (average number of edges) for each node is pN, where N is the number of nodes (if self-connection is not allowed this is p (N − 1) ≈ pN for N> > 1 ). The degree distribution is P(d) = C(N;d) p d (1 − p)N d where C(N;d) is the number of different ways one can choose d objects out of N, regardless of order. The binomial distribution reduces, for small p and large N, to a Poissonian, with the mean and variance equal to pN.

The K-core decomposition of a graph is obtained in the following way: Throw out the disconnected nodes. Find the set of all nodes that have only one edge S1. Now remove all those single edges. Add to S1 all those nodes which now end up being singly connected and iterate until no singly connected nodes remain. Put in S2 all those nodes which now have two edges. Cut those edges and add to S2 all those nodes which have only two edges remaining as a result; iterate until you do not encounter any nodes with two edges any more. Repeat for 3, 4, …K edges, until you exhaust all the nodes. In this way, the graph is partitioned into K sets (‘shells’) that are successively more highly connected the larger K is. The number of such shells is an indicator of the complexity of the network.

The neighbourhood of a node (say node i) is defined as all those nodes to which i is directly connected, i.e. which are only one edge away from i. The clustering coefficient of a node i, namely c i , is defined as the probability that the nodes in the neighbourhood of i are connected between themselves pairwise, and can be computed as the number of edges connecting the neighbours of i to each other, normalized by the number of distinct pairs of nodes in the neighbourhood of i. The clustering coefficient of a graph is found by averaging the clustering coefficients of the individual nodes. Clearly one can also define clustering coefficients averaged over different subsets of nodes, e.g. c(d) is the clustering coefficient averaged over all nodes having degree d.

The rich-club coefficient r(d) is defined as the probability that nodes having degree larger than k are connected pairwise between each other. It can be computed as the number of edges running between nodes with degree > d, normalized by the total number of distinct pairs of nodes with degree d.

The degree-degree correlation function is the correlation between the degrees of a node and the degrees of its neighbours. It is defined as d nn = ∑ d' p(d | d') where the sum is over d' and p(d | d') is the conditional probability that a neighbour of a d-degree node has degree d'.

2.3 Appendix 2.3 Weight distribution of the miRNA-miRNA co-regulation network

Let A be the M×N adjacency matrix of a bi-partite graph, where there are two types of nodes, let us say a-type nodes and b-type nodes (as in the case of miRNA and mRNA), and a-type nodes can only be directly connected to the b-type nodes.

Then W = AAT (where the superscript T denotes ‘transpose’) can be thought of as a ‘weighted’ adjacency matrix of the network of a-nodes, with the interactions being mediated by the b-nodes. The weight of an interaction is equal to the number of different b-nodes nodes via which two a-nodes are connected to each other.

The elements of W (the weighted M × M adjacency matrix of the miRNA-miRNA network, in our case) are given by \( {w}_{ij}={\left(\mathbf{A}\kern0.5em {\mathbf{A}}^T\right)}_{ij}={\displaystyle \sum_{k=1}^N{A}_{ik}{A}_{jk}} \).

The off-diagonal elements are nonzero if the nodes i and j (both of type a) are connected to each other by at least one path consisting of two successive edges, leading from the ith type a-node to some type b-node and then back to the jth a-node (recall that nodes of the same type are not connected to each other in the miRNA-mRNA network). The weight w ij is in fact equal to the number of such parallel two-step paths, between the ith and jth miRNAs. The diagonal elements are equal to twice the degree of the a-nodes in the original a-b network (the two-step paths in this case return to the ith a-type node from which they issued). The frequency f(w) of any given weight w among the off-diagonal elements of the matrix W, normalized by (1/2) times the number of off diagonal elements, gives the probability distribution function for the weights, P(w), viz.,

$$ P(w)=\frac{2f(w)}{M\left(M-1\right)}. $$

Here the factor of (1/2) comes from the fact that the number of off-diagonal elements is equal to twice the number of a-type pairs of nodes; M is the number of a-type nodes.

Appendix 3

3.1 Appendix 3.1 Sequence matching probabilities between random sequences

Let us consider sequences made up of characters drawn with equal probabilities (1/r) from an r-letter alphabet. The probability of one-to-one sequence matching (or Watson–Crick base-pairing) of a random sequence of length l' at least once within a longer random sequence of length ll' is

$$ p\left(l,l^{\prime}\right)=1-{{\left(1-{r}^{-l}\right)}^{l\prime}}^{-l+1}. $$

to a very good approximation (Mungan et al. 2005 and references therein). It should be noted that for l' > > l, p(l,l') ~ l' /r l .

If we make the simplifying assumption of equal probabilities of occurrence for all the nucleotides in these sequences, this relation can be used in two problems we have dealt with in this article: (i) for the probability of encountering the seven-nucleotide-long ‘seed’ sequence of the miRNA within the 3′ untranslated region of an mRNA, and (ii) the probability of encountering the consensus sequence for a transcription factor in the promoter region of a gene.

These considerations can easily be generalized to the case where the letters occur with different probabilities, and where the sequences also display nearest neighbour correlations between their elements. If the probabilities of occurrence of the different ‘letters’ (nucleotides) – in either the short (e.g. seed) or long (e.g. in the UTR) sequences – are appreciably different from each other, one has to replace r −l in p(l,l' ) by the appropriate product of the probabilities for the individual letters making up the seed region, within the zeroth order Markov approximation, or the product of the conditional probabilities within the first order Markov approximation (appendix 4).

3.2 Appendix 3.2 Computing the in-degree distribution for the miRNA-mRNA target network

Let us consider the example of the miRNA-mRNA target network. To compute the in-degree distribution (figure 2), we need a ‘binding probability’, i.e. probability that the seed sequences of a miRNA occurs at least once in the UTR of any given mRNA. The seed sequence on the miRNA is taken to have a fixed length (l = 7).

For sequences consisting of symbols drawn independently and with equal probabilities from a set of cardinality r, the probability to encounter any specific sequence of length l is simply given by (1/r)l. In the case of the nucleotide alphabet A, T, C, G, clearly r = 4.

The distribution of the number of edges terminating on any mRNA with a UTR of length l' is given by the binomial distribution,

$$ {P}_{l\prime }(d)=C\left(M;d\right)p{\left(7,l\prime \right)}^d\left[1-p{\left(7,l\prime \right)}^{M-d}\right] $$

for a set of M miRNAs. The total in-degree distribution is then simply given by

$$ P(d)={\displaystyle \sum {P}_{l\prime }}(d)g\left(l^{\prime}\right) $$

where g(l') is the length distribution of the UTRs and the sum is over all allowed l'. The empirical, point for point length distribution of the UTRs was obtained from the databases (Grillo et al. 2010). We found that we could fit this distribution with a smooth function

$$ g(k)=\mathrm{N}{k}^{\psi }{e}^{-k/\kappa }, $$

where the normalization factor is given by N = Γ(ψ + 1)κ ( ψ+1 ), Γ is the gamma function, and the values for the parameters ψ and κ for different species can be found in the following table.

Fitting parameters and first moments of UTR length distributions of the mRNAs of four organisms considered in section 3.1

Parameters

H. sapiens

B. taurus

X. tropicalis

C. elegans

κ

515

485

490

202

ψ

0.14127

0.261554

0.254772

0.253727

<k>

874

739

731

338

The degree distribution which is computed here for the miRNA-mRNa network is therefore a weighted superposition of different Poisson distributions (classical random networks – see appendix 2.2), with different connection probabilities p(7,l′), depending on the different UTR lengths (l′) present.

Appendix 4 Generating sequences via zeroth and first order Markov processes

Here we provide a working definition of a Markov process which generates a Markov sequence or Markov chain. More formal and correct definitions can be found in standard textbooks on probability theory.

A sequence (when we say a sequence we mean an ordered sequence) x 1, x 2 … x N is called Markovian if the ith element in the sequence, x i depends at most on the values of a finite number n of immediately preceding elements, i.e. {x i − n , x i − n + 1 … x i − 1}. The Markov sequence does not depend on the whole ‘history’ of the process.

In a Markov sequence of order 0, each element is chosen independently, according to a fixed probability distribution {p I }, from the set of allowed elements (or ‘values’) X I with I = 1,2, … M. Thus, the sequence has no memory of the preceding part of the sequence at all. Thus, the probability of encountering any particular subsequence of length l = 3, e.g. (X 15 X 3 X 7), starting from any given position in the sequence is simply given by the probability of the product of the elements which appear, namely p 15 p 3 p 7. In order to model a naturally occurring sequence of length N, one may estimate the probabilities p I of the different characters from their frequencies N I of occurrence, so that

$$ {p}_I\approx {N}_I/N. $$

For a Markov sequence of order 1, each subsequent element x i+1 does depend on the preceding one, i.e. x i . One then needs the set of M × M conditional probabilities, p(X I |X J ) = p(X I ,X J )/p(X I ). On the left-hand side we have the conditional probability for X J to appear, given that the preceding element is X I . The equality is known as the Bayesian identity. On the right-hand side we have the joint probability for the elements X I and X J , divided by the probability of the element X J . Then, the subsequence used in the example above has the probability p(X 15|X 3)p(X 3|X 7) to be encountered after the element X 15.

In order to model the first order Markov process from which a sequence may have originated, one may once again estimate the conditional probabilities p(X I |X J ) from the conditional frequencies N IJ via p(X I ,X J ) ≈ N IJ /(N − 1), where N IJ counts all 2-element subsequences X I X J in the entire sequence, and divides by the total number of 2-element subsequences, which is N−1 ≈ N.

Tables of probabilities used for generating zeroth and first order Markov series for miRNA seed and mRNA 3′ UTR sequences have been provided as supplementary material.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Arpağ, G., Erzan, A. An ensemble approach to the evolution of complex systems. J Biosci 39, 259–280 (2014). https://doi.org/10.1007/s12038-013-9394-8

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12038-013-9394-8

Keywords

Navigation